Carbonate Acidizing - A Review On Influencing Parameters of Wormholes Formation (2023)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Journal of Petroleum Science and Engineering 220 (2023) 111168

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Carbonate acidizing – A review on influencing parameters of


wormholes formation
Cláudio Regis dos Santos Lucas a, Jair Rodrigues Neyra a, b, Elayne Andrade Araújo c,
Daniel Nobre Nunes da Silva a, b, Mateus Alves Lima a, David Anderson Miranda Ribeiro a,
Pedro Tupã Pandava Aum a, b, *
a
Federal University of Pará (UFPA), Petroleum Science and Engineering Lab (LCPetro), Salinópolis Campus, 68721-000, Salinópolis, Brazil
b
Federal University of Pará (UFPA), Postgraduate Program of Chemical Engineering (PPGEQ), 66075-110, Belém, Brazil
c
Federal University of Rio Grande do Norte (UFRN), Postgraduate Program in Chemical Engineering (PPGEQ), Lagoa Nova, 59078-970, Natal, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: Acidizing is a well-stimulation operation that consists of injecting a reactive fluid into the rock formation. When
Wormholes in carbonate rocks, the dissolutions create conductivity channels called wormholes. The pattern formed depends
Acidizing on the flow rate, thermodynamic conditions, and several rock-fluid interaction parameters. Despite acidizing
Carbonate
operations being well-known, several factors or conditions affecting wormhole formation are not thoroughly
PVbt
reactive flow
tested in the laboratory. We observe a difficulty in the literature to summarize the main aspects involved in
wormhole formation. At the same time, understanding how each parameter could affect the wormholing process
can help to optimize the acidizing design, maximizing the financial return. Therefore, this review article dis­
cusses the main studies about the parameters affecting the wormhole’s formation: acid concentration, reaction
rate, flow rate, temperature, core sample dimension, and heterogeneity. The main idea here is to provide a
resume of the most relevant works founds in our literature review and a reference base for researchers interested
in carbonate acidizing. The pore-volume-to-break-thought (PVbt) plotted as a function of the flow rate is the
most common approach to evaluate the dissolution pattern observed for each reactive fluid-rock combination.
However, PVbt should be seen more comprehensively as a consequence of advection-diffusion-reaction balance.
Other essential aspects that need to be considered to obtain a significant representation of the PVbt plots are
sample geometry and the initial rock saturation.

1. Introduction operational parameters. One of the project design challenges is having


rock-fluid information, especially in new projects. The lack of knowl­
Acidizing is a stimulation technic widely applied to enhance oil edge of rock-fluid interactions or treatment application conditions can
production. It consists of injecting a reactive mixture, normally con­ lead to treatment failures and, in some scenarios, even formation
taining HCl, into the reservoir at a pressure below the rock fracture damage.
strength. The acid reaches a few meters in the near-wellbore region, The HCl (Hydrochloric Acid) is the most common acid used, mainly
dissolving or dispersing the damage. The rock matrix reacts with the due to its price, availability on the market, and high reactivity with
acid in carbonates formations, forming conductive channels called carbonate formations. The HCl is generally supplied with 32–36% wt.
wormholes (Kalfayan, 2008). and is usually diluted, on the formulation, until 15% wt. Organic acids
The wormhole formation is complex; it involves reaction-diffusion- are also used to stimulate carbonate formations. Acetic and formic are
advection, thermodynamics properties, and rock-fluid interaction. the most common and can form wormholes. In addition, organic acids
Several advances have been made to improve the understanding of the are less reactive than HCl, offering an alternative in high-temperature
phenomena. Realizing the dominant parameters in wormhole formation scenarios where the corrosion rate increases considerably for HCl,
helps optimize the acidizing formulations and set up the correct especially in cases with restricted pump availability due to horsepower

* Corresponding author. Federal University of Pará (UFPA), Petroleum Science and Engineering Lab (LCPetro), Salinópolis Campus, 68721-000, Salinópolis, Brazil.
E-mail address: pedroaum@ufpa.br (P.T. Pandava Aum).

https://doi.org/10.1016/j.petrol.2022.111168
Received 23 April 2022; Received in revised form 2 October 2022; Accepted 25 October 2022
Available online 28 October 2022
0920-4105/© 2022 Elsevier B.V. All rights reserved.
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

or safety limits (Ali et al., 2016). ( )


Another way to control the acid-rock reaction is to promote a qm,A = km CA,bulk − CA,solid surface Eq. 1
physical delay to the acid diffusion through fluid, such as using viscous Being qm,A the mass flux, km is the diffusion coefficient, CA,bulk , and
fluids or trapping the HCl in a dispersed phase of emulsions (Rae and di CA,solid surface are the acid concentrations in bulk and at the solid surface,
Lullo, 2003). In addition, the reaction control avoids the acid being respectively. The adsorption and desorption processes are very fast and
completely consumed near the injection point, allowing deeper generally not modeled, considering that all acid is adsorbed instanta­
treatments. neously on the surface and is available for reaction. The products are
As we stated before, in carbonate acidizing, the acid reacts, dis­ immediately desorbed when formed. Indeed, this approach is used for
solving the rock, and the wormholes are formed. The optimal condition the HCl-carbonate reaction. Therefore, the reaction could be considered
is when the minimum acid volume achieves deeper reservoir regions, irreversible and first order, as described in Eq. (2).
assuring a better well-reservoir connection. Wormhole formation is a
phenomenon wholly related to rock-fluid interaction that depends on an rHCl = kr CHcl,solid surface Eq. 2
equilibrium of the advection-diffusion-reaction process. Additionally,
various variables such as thermodynamic circumstances, rock satura­ Once the reaction completely consumes the flux, we can state that
tion, effective pore fluid velocity, heterogeneity of the rock, and tortu­ rHCl = qm,HCl . Therefore, we can extract the concentration at the bound­
osity might alter this equilibrium (Ali et al., 2016; Kalfayan, 2008; ary by combining equations.
Theresa Sibarani et al., 2018; Zakaria et al., 2015a; Ziauddin and Bize, qm,HCl = rHCl Eq. 3
2007).
( )
We found several works in the literature studying the wormhole km CHCl,bulk − CHCl,solid surface = kr CHCl,solid surface Eq. 4
formation process from specific perspectives. This review article aims to
discuss the essential factors that affect wormhole formation and pro­ Isolating CHCl,solid surface we find:
vides readers with a broad view of the mechanism and elements km
involved in carbonate acidizing. The simulations performed to illustrate CHCl,solid surface = CHCl,bulk Eq. 5
(kr + km )
the effect discussed in the following subjects were accomplished with in-
house software developed by Science and Petroleum Engineering Lab­ Now we can re-write the reaction rate as a function of the concen­
oratory (LCPetro/UFPA), based on the mechanistic model proposed by tration of the bulk solution, as follows:
Ali and Ziauddin (2020). km
rHCl = kr CHCl,solid surface = kr CHCl,bulk Eq. 6
(kr + km )
2. Chemical of Rock-HCl
kr km
The wormhole formation happens due to the interaction of an rHCl = CHCl,bulk Eq. 7
(kr + km )
reactive fluid flowing in a carbonate rock porous media. The reaction is
heterogeneous, once involves two phases, the reactive fluid, normally an rHCl = keffective CHCl,bulk Eq. 8
acid, and the carbonate rock that is the solid phase. The reaction occours
when the acid reaches the rock surface. In practicality, this implies that How we are specifying the reactions with HCl, we can recall Eq. (1),
the acid needs to be transported until the rock surface to reacts. Fig. 1 where we can also observe that once the acid is consumed very fast, the
ilustrate the main steps of the reactive flow, as follow: (1) the acid is concentration on the surface is near-zero or at least is
firstly transported to the boundary layer by advection; (2) then diffuses CHCl,boundary ≫CHCl,solid . Therefore, we can state that
on the boundary layer until it reaches the rock surface; (3) the reaction ( ) ( )
qm,HCl = rHCl = km CHCl,bulk − CHCl,solid surface ∼ km CHCl,bulk Eq. 9
occours and the acid is consumed in the rock surface and products are
formed; (4) products dessorb from the rock surface and diffuse trough It is important to remember that these expressions are derived for the
the boundary layer; and finally (5) are carried by advection due to the reaction HCl-Calcite, just to illustrate the aspects involved. If the rock or
macroscopic fluid movement. acid are changed, the equations, and the considerations used need to be
The subscript “A” indicates the reactive component; for example, the changed.
HCl, “M” represents the mineral, and “P” represents the products Before discussing in detail these types of structures formed, called
formed. wormholes, we will introduce the experimental approach used to
The diffusion process at the boundary layer can be described by Eq. perform the reactive flow experiments with focus on carbonate
(1): acidizing.

3. PVbt experimental determination

One of the main carbonate acidizing experiments performed in the


laboratory is coreflooding. Field conditions are applied to the core
sample and the acid is injected to evaluate the reactive flow. The ex­
periments to obtain PVbt are carried out in Hassler single core cell
systems, where it is possible to confine the rock samples and carry out
the injection of water, brine, and acid systems. An essential point in
carrying out this type of experiment is that the dissolution reaction
generates CO2, so it is necessary to maintain backpressure above 1100
psi to ensure that the flow does not become biphasic into the rock
(Kumar et al., 2021; Shukla et al., 2006; Wang et al., 1993). What could
interfere with the mass transfer process and the manometer readings,
consequently, compromising the experiment results. Another critical
point is that all the acid-wetted parts of the equipment must be made of
Hastelloy or other material with equivalent resistance to avoid leaks and
Fig. 1. Heterogeneous reaction mechanism.

2
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

minimize the possibility of accidents. the penetration of acid into the formation for the same amount of acid.
Some coreflooding systems are dual cores, which aim to evaluate This concept is represented in Fig. 6, where five main types of structures
divergent formulations systems. The experiment is very similar; rock are observed. We will discuss each dissolution pattern later in the dis­
samples with contrasting permeabilities are placed in a parallel config­ cussion of flow rate.
uration. The ability of chemical systems to diverge the flow from higher The point here is to understand that different injection conditions or
to lower permeability regions, which naturally would not occur, is rock-fluid interactions will provide different dissolution patterns.
generally studied. Fig. 2 shows conventional equipment of coreflooding Correlating this with carbonate acidizing reflects how deep the acid can
used for carbonate acidizing, and Fig. 3 shows a schematic set up. penetrate the reservoir, the type of conductivity structure formed in the
Pore volumes to breakthrough (PVbt) is defined as the pore volumes rock, and the volume of acid injected to form the structure.
of acid, or reactive fluid, necessary to allow the channel formed, during
the dissolution process, breakthrough the core sample. The experiment 4. Acid concentration
is conducted by confining a rock saturated with distilled water or brine
(in general, we will discuss this point later). After stabilization, pressures Carbonate minerals react completely with excess hydrochloric acid
are taken at the same injection rate, in order to calculate the perme­ (HCl), producing water, carbon dioxide, and water-soluble salts. The
ability, and then acid injection into the rock sample begins. supplied acid concentrations vary between 32 and 36% wt. and are
The reactive fluid then percolates through the porous medium, and diluted in the range of 5–28% wt. for each application condition. Since
pressures across the sample are measured. When the injected fluid reacts the first acidification operations, HCl has remained the primary acid
with the rock, dissolving it, the porous medium deteriorates, thus used to stimulate carbonate formations, with 15% wt. being the standard
increasing the porosity of the medium and forming high permeability concentration used in the field. It is intuitive to think that increasing acid
channels. When crossing the sample, the differential pressure between concentration will reduce PVbt. Once more acid is available in the me­
the inlet and outlet of the rock sample is practically zero, as there is no dium, faster will be the dissolution, and consecutively, faster will be the
resistance to flow. The acid pumping is then stopped. This point is advancement of the wormhole in the rock. Furthermore, that is what is
generally known as the breakthrough point. This experiment generates a observed. However, an important aspect to discuss is the behavior as the
characteristic differential pressure behavior, shown in Fig. 4. As soon as acid concentration rises more and more.
the acid reaches the rock formation (red line), the pressure drops The graph in Fig. 7 shows how PVbt shifts with increasing acid
slightly. Until reaching a minimum value where it stabilizes (blue line), concentration. These curves were obtained with an in-house simulator
at this point, the channel has already crossed the sample. of LCPetro, obtained through the model described by Ali and Ziauddin
The remaining pressure reading is due to the pressure drop of the (2020). The simulation was made considering core dimensions of 1.5 in
fluid crossing the formed channel as a free medium, so depending on the of diameter and 6 in of length. The flow rate varied from 0.5 up to 20
sample length, flow rate, and equipment characteristics, this value must mL/min in 20 steps. The temperature was fixed at 45 ◦ C.
be close to zero. The volume of reactive fluid used to allow the wormhole A priori, as we increase the concentration, we shift the PVbt curve
breakthrough of the sample, normalized by the initial rock sample pore downwards, as shown in Fig. 7, where we noticed that as we increase the
volume, is called pore volume to breakthrough (PVbt). concentration of HCl, the curves tend to overlap. Therefore, we observe
PVbt is dimensionless and could be calculated using Eq. (10), where: progressively low changes in optimal PVbt as we increase to concen­
Vbt is the volume of acid used to breakthrough the rock sample; Vap is the trations above 15% wt. This indicates that even with high acid avail­
apparent volume of the rock; φ is the porosity. ability, there is a limit promoted by the reaction process.
Still, in Fig. 7, the optimal points for each PVbt curve are shown by
Vbt
PVBt = Eq. 10 yellow points. The optimal points were obtained by successive approx­
Vap ⋅φ
imation in order to determine the lowest PVbt value in the curve.
The PVbt is obtained for different flow rates, so a curve is set up, We can observe that the optimal point is displaced to the right once
called the PVbt curve shown in Fig. 5. The PVbt will be affected for each the concentration increases. However, when the optimal point is dis­
fluid type, rock, and injection condition. Thus, changing the rock type placed to the right, we need higher flow rate values, which, from an
will change the curve, as shown in the plots in Fig. 5. The “f” reported in operational aspect, could be unreachable in scenarios of limited horse­
each plot represents the respective flowing fraction, a concept to char­ power and extremely low permeabilities.
acterize the heterogeneity of the rock that we will discuss further ahead. Some authors outlined that the properties of high HCl concentrations
As previously mentioned, the PVbt approach will be linked to the solutions are very different from low concentrations, and the effect of
relationship between the amount of acid to be injected and the depth using high concentrations, in the range of 20 up to 28 % wt., could be
reached by the conductivity channel formed in the dissolution process beneficial for some scenarios (Harris et al., 1966; Kalfayan, 2008).
(wormholes). This means that a priori, the lower the PVbt, the greater Another concern with increasing the concentration is that the

Fig. 2. A conventional system for reactive flow is shown, and the schematic drawing is in Fig. 3.

3
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

Fig. 3. Schematic of set up equipment for reactive flow experiments.

Fig. 5. Experimental PVbt (Zakaria et al., 2015a).

literature (Fredd and Fogler, 1998; Furui et al., 2010; Wang et al., 1993).
Their interests involved the development of a model that would bring
the optimal rates for acidification coreflooding experiments, which
Fig. 4. Characteristic curve of reactive axial flow experiment. could be applied in the field. According to their results, a higher acid
concentration results in a higher optimal acid flow rate and progres­
reaction rate increases to a limit, where the formed products start to sively lower optimal PVbt values. Fig. 8 shows the simulations per­
inhibit higher rates. However, defining the best concentration is still a formed by Xue et al. (2018) and the experimental results obtained by
relatively complex, as it depends on many variables. Wang et al. (1993) when increasing the acid concentration.
According to an experimental and theoretical study by Dong et al. An essential point in this discussion is that injected acid contains
(2016), the optimal injection rate of acid treatment depends on factors additives or even it is a component of a fluid system. Thus, its concen­
such as rock lithology, acid concentration, temperature, and rock pore tration will not necessarily be effective in the rock-fluid reaction since it
size distribution. The experiments performed by Dong et al. (2016) interacts with the continuous medium (fluid bulk) and with other ad­
involved samples 1.5 inches in diameter and 8 inches in length for ditives. Therefore, looking only at concentration, in general, is not
carbonate formations of the following rock types: Indiana limestone, enough. Instead, we need a broader analysis.
Laoux limestone, and Kansas chalk, with characteristics reported in the

4
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

5. Retarded acid system

As discussed in the previous section, reactivity directly affects


channel development in porous carbonate rocks. Strong acids tend to be
wholly consumed in the near wellbore region, which limits the treat­
ment penetration, forms thicker channels, and promotes lower worm­
hole density. Furthermore, as observed in Fig. 7, the higher acid power
tends to favor the shift of optimum PVbt to higher flow rates, making it
technically and financially challenging to achieve in the field, especially
in tight formations due to pump operational limits.
Retarded acid systems are formulations developed to achieve deeper
rock formation before the acid is wholly consumed, extending the
reservoir contact compared with regular HCl (Abdrazakov et al., 2018;
al Moajil et al., 2020; Bernadiner et al., 1991; Hoefner and Fogler, 1987;
Khan et al., 2021; Maheshwari et al., 2013; Paccaloni et al., 1988).
Several formulations were developed with the intent of reducing the
reactivity or reducing the acid diffusion. Many of these solutions were
adopted by oil services companies with specific formulations. Retarded
acid systems generally include organic acids, chelates, emulsified acid,
and chemical retarders (al Moajil et al., 2020, 2019; Bernadiner et al.,
Fig. 6. Dissolution structures formed during flow and reaction in carbonate 1991; Crowe et al., 1990; Hongping et al., 2022; Huang et al., 2003;
porous media, and corresponding pore volumes of fluid required for break­ Shafiq and Mahmud, 2017). Other systems, based on surfactants, poly­
through (Fredd et al., 2017).
mers, gelling agents, and foamed acid, also have a retarded effect
(Arruda et al., 2022). Gelling acid systems can reduce acid diffusion but
with less intensity of emulsions. Therefore, they are generally preferred
as diverting acid systems or fracking applications.
Organic acids, such as Formic and acetic, emerged as the first al­
ternatives to avoid the high reactivity in acidizing caused by conven­
tional HCl treatments. These acids present low H+ dissociation,
allowing a more controlled acid-rock reaction. Since then, several other
systems based on organic acids have been developed. Under the same
conditions, the optimum flow rate for weaker acids is significantly lower
than that found for strong acids. This compensates for the reduction in
reactivity during the flow in porous carbonate media (Akanni and
Nasr-El-Din, 2016; Bernadiner et al., 1991; Buijse et al., 2003).
Fredd and Fogler (1997, 1996) use chelate agents as acidizing
stimulating fluid. Their result showed that the ethylenediaminetetra­
acetic acid (EDTA) could form wormholes in carbonates at low flow
rates without face dissolution. Further studies also demonstrate the ef­
ficiency of other chelates such as hydroxy ethylethylenediaminetetra­
acetic acid (HEDTA), diethylenetriaminepentaacetic acid (DTPA), and
L-glutamic acid-N, N-diacetic acid (GLDA). One interesting point here is
that chelates’ mechanisms for wormhole formation in carbonates consist
of sequestering the calcium ions from the carbonate rock matrix. The use
of chelates also helps reduce precipitation issues (Adenuga et al., 2013;
Fig. 7. PVbt plots for different HCl concentrations. Ali et al., 2008; Beuterbaugh et al., 2015; de Antuñano et al., 2015;
Fredd and Fogler, 1996, 1997; Hassan et al., 2020; Huang et al., 2003;
Husen et al., 2002; Khan et al., 2021; Mahmoud et al., 2017; Parkinson

Fig. 8. PVbt plots for different HCl concentrations for numerical simulation (Xue et al., 2018), and experimental result (Wang et al., 1993).

5
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

et al., 2010; Reyes et al., 2013; Tariq et al., 2021). that at a low flow rate using in-situ foam, the PVbt value decreases
Emulsified systems applied to carbonate acidizing consist of acid significantly and avoid face dissolution pattern in limestone under 0.5
dispersion in an oily continuous phase. The primary mechanism is to cm3/min. Fig. 9 (a) shows the dissolution pattern obtained compared to
trap the acid in the dispersed micelle, drastically reducing the diffusion conventional treatment. When considering dolostone, typically reaction
process of the acid in the continuous phase (Al-Mutairi et al., 2012; rate is limited to below 122◦ F. The in-situ foam generated allowed to
Bazin and Abdulahad, 1999; Chacon and Pournik, 2022; Ge et al., 2020; reach a PVbt and a well-formed wormhole, like limestone, as shown in
Oliveira et al., 2013; Pandya and Wadekar, 2013; Siddiqui et al., 2006). Fig. 9 (b), in contrast to no-channel formed in conventional treatments
Emulsions are widely used in field operations, and the services com­ at flowrates of 0.25 cm3/min.
panies make several formulations available.
Microemulsions are also reported as retarded acid systems, with the 6. Flow rate
advantage of present droplet sizes of 1 up to 100 nm, and are thermo­
dynamically stable (Dantas et al., 2019). Furthermore, due to his The injection flow rate is one of the critical factors in the acidifica­
reduced droplet size, microemulsions exhibit considerably lower vis­ tion of matrices. It directly influences the dissolution patterns formed in
cosity than regular emulsions. (Andreasson et al., 1993; Aum et al., the rocks arising from the reactions with the acid. As previously illus­
2016, 2021; Carvalho et al., 2019; de Castro Dantas et al., 2019; Hoefner trated, the PVbt curve is obtained by varying the flow rate and fixing all
and Fogler, 1985; Yarveicy et al., 2018). Nevertheless, high surfactant other parameters such as core length, core diameter, temperature, fluid,
concentrations are typically required, reducing cost-effectiveness. etc. Fig. 10 shows the different patterns observed varying the injection
Chemical retarders, or inhibitors, focus their mechanism on reducing rate condition. Five main dissolution patterns are observed: (1) facial
the reaction rate between HCl and carbonate rocks by creating a pro­ dissolution; (2) conical wormhole; (3) dominant wormhole; (4) ramified
tective film over the surface rock (Hoefner and Fogler, 1987; Hongping wormhole; and (5) uniform dissolution. The first dissolution pattern
et al., 2022). In this way, Oxyalkylated alcohols, fatty amines, ampho­
teric/anionic surfactants, and sulfonate/phosphonate compounds form
a stabilized thin layer on the rock surface, thus creating a physical
barrier (Crowe et al., 1990; Wanderley Neto et al., 2021). By another
route, water-wetting surfactants decrease the reaction rate by striping
the oil saturating from the carbonate rock surface. Recently, Hongping
et al. (2022) reported using two synthesized acidizing retarders. The
first was prepared with acrylamide and allyl polyethylene glycol. The
second retarded system was the same first formulation added with
octadecyl dimethyl allyl ammonium chloride. Authors found that the
increase in the hydrophobic chain increases the adsorption film thick­
ness, thus raising the retardation power.
Foamed acids are generally used as divergent agents but also exhibit
retarded properties. Foams prevent acid from spending outside the
primary dissolution channel and, as a result, reduce branching. This
effect results from reducing liquid saturation, thus decreasing effective
permeability to liquid in secondary channels (Bernadiner et al., 1991; Li
et al., 2008). According to Li et al. (2008), at high capillary pressure, the
apparent yield stress of foam increases, and many bubbles are trapped in
those pores. Bernadiner et al. (1991) performed experiments using HCl
at a concentration of 3 N and dodecylbenzene sulfonic acid in Texas
cream chalk and dolostone cores measuring 1.5 in (diameter) and 1–4 in Fig. 10. General plot of PVbt as a function of injection rate (pictures from:
(length). Although the lack of a comprehensive PVbt plot, they found Fredd et al., 2017).

Fig. 9. Effect of in-situ generated foam on PVbt in (a) Limestone and (b) Dolostone (Adapted from: Bernadiner et al., 1991).

6
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

occurs at a very low injection rate called compact or face dissolution resulting in a wormhole with a certain degree of tortuosity. Therefore,
(Wang et al., 1993). The main characteristic is the dissolution concen­ the main channel length will be greater than the length of the core. Thus,
trated in the injection core face. The wormhole cannot be formed once there seems to be a relationship between the main channel tortuosity
the advective transport is minimal and the reaction is intense. Face and the ramifications formed due to acid leakage from the main channel.
dissolution is critical to obtain in the laboratory because the time is Some authors have observed that longer branches around the main
extensive once we have a very low flow rate and core crushing is very channel result in a less tortuosity main wormhole. The expenditure of
likely to happen. Also, in numerical simulations, this pattern presents a acid in these branches makes propagation less tortuous through the core
long time to run entirely, and usually, simulations are stopped after a , which is not the optimal scenario (Yoo et al., 2022). The optimal
defined simulation time. (idealized) scenario would be a narrow channel without tortuosity,
In the sequence, slightly increasing the flow rate above the face connecting the input and output of the plug.
dissolution regime, we have the second pattern, which is called a conical Another effect linked with flow velocity could be observed in the
wormhole. We also observe a dissolution in the injection face, but less works of Izgec et al. (2009, 2010), discussed in more detail in the section
intense than in the face dissolution pattern, and we observe the forma­ on the impact of heterogeneity. The authors analyzed the presence of
tion of one main channel. Due to low velocity, the acid is mainly vugs during the acidizing experiments. Their results showed that the
consumed at the beginning, and the acid concentration inside the wormhole tends to follow regions with higher concentrations of vugs
channel decreases with the distance from the face of the injection. since they provide local pressure gradients that facilitate the acid flow
Therefore, a conical shape structure is formed. The dominant wormhole path and consequent wormhole propagation. Although this effect is
is the preferred path once it requires the lowest acid volume for the related to pore heterogeneity, it is closely linked to effective flow ve­
sample’s channel breakthrough. This pattern is characterized by forming locities in the porous medium.
a narrow main channel with a regular diameter throughout the entire Thus, variations in petrophysical properties, permeability, and
sample. The broad base of the cone is not more present. In this pattern, porosity change the morphology of the wormhole and its preferential
most of the acid is converted to advance the wormhole; therefore, few path.
branches are observed, and the rock volume dissolved is minimum.
Above the optimal flow rate, we have the dissolution pattern called a 7. Combining reaction and advection – dimensionless approach
ramified wormhole, which presents several branches formed due to acid
leaks off from the main channel due to the limited advance of the acid As previously discussed, if we change the reactive system or the rock,
for advection, a consequence of the pressure pattern into the porous the PVbt curve changes. An alternative approach that can consider both
medium. In this pattern, acid consumption is higher than optimal; is through dimensionless numbers. Currently, two main dimensionless
consequently, more rock is consumed. When the flow rate increases, the parameters are used to evaluate these channels’ formation: the Peclet
leak-off also increases. Therefore, even more, branches are formed. and the Damköhler number (Economides and Nolte, 2000; Hoefner and
When the density of branches is too high, the main channel is not more Fogler, 1987, 1988). Damkhöler number (Da) represents the ratio of
trackable. At this point, we have the pattern called “uniform dissolu­ contributions of dissolution rate and the acid injection rate as following
tion”. One interesting point is that high pressure is verified, and some­ ( )
times it is difficult to achieve the breakthrough without fracturing the L
Eq. 11
2
Da = a D3e ⋅
rock. q
Therefore, at extreme flow rates (low and high) the wormholing
where L (cm) represents the core sample length, De is the effective
formation is inefficient (Wang et al., 1993).
diffusion coefficient (cm2/s), q is the flow rate (cm3/s), and “a” is a
Observing the dissolution patterns presented above, we realized that
constant that varies with the rock sample. Optimum wormhole forma­
the reactive flow and the wormhole formation are definitely impacted
tion occurs at Damköhler values near 0.29 (Fredd and Fogler, 1998). In a
by the advective transport represented by the injection flow. The exis­
different approach, the Peclet number is defined as the ratio of the
tence of an optimal point indicates that there is a balance between the
convective contribution to the diffusive flux. This dimensionless con­
reaction and the transport processes leading to an efficient propagation
siders only mass transport as a limiting factor for the process. The range
of the wormhole (Fredd and Fogler, 1999). Therefore, in the flow in the
of the Peclet number that favors the formation of wormholes is between
porous medium, the size of the pores, their distribution, and the geo­
10− 3 and 10− 2, and the calculation can be done through Eq. (12).
metric heterogeneities in the rock matrix are factors that will affect the
√̅̅̅
effective flow velocity. Rocks can have a high diversity of pores with q⋅ k
different sizes and shapes, which adds complexity to the flow path. Pe = Eq. 12
A⋅De
Therefore, the heterogeneity of the rock matrix will also contribute to
our understanding of how the wormhole will propagate. Where: q is the flow rate (cm3/s); k is the permeability (cm2); A =
Another important observation regarding the flow distribution is transversal area (cm2), and De is the effective diffusivity (cm2/s).
that we also have losses to the radial direction. Thus, Huang et al. (2000) Daccord et al. (1993) studied porous media dissolution under reac­
observed that the loss of fluid through the walls of the main channel is tive flow. They concluded that the experimental results could be situated
one of the main factors that impact the efficiency of wormhole in a phase diagram with delimited regions (or domains) of dissolution
propagation. pattern. The representation was obtained by plotting a dimensionless
The results presented by Ahoulou et al. (2020) showed that during kinetic number (ki) versus the Peclet Number. The kinetic number is
the acid flow, there is a transition between the dissolved region related to Peclét and Damköhler by Eq. 13
(wormhole) and the undissolved region (porous medium with no acid
ki = Pe⋅Da
attack), indicating that there is a porosity gradient in this transition
region. This porosity gradient would facilitate the creation of branches Their results show a straightforward transition for different flow
from acid loss from the main channel. (Ahoulou et al., 2020). regimes.
Thus, the amount of acid that reaches the tip of the wormhole during In this context, Fredd and Fogler (1998) developed a comprehensive
its propagation depends on the fluid losses in the wormhole walls study about the Damköhler number, which intends to incorporate all the
(Huang et al., 2000). Therefore, the acid velocity at the wormhole’s tip physics of the reaction process. Once the acid or reactive system needs
differs from the other parts of the channel. Furthermore, considering the diffuse from the solution into the pore wall, the wall reaction occurs, and
dominant wormhole, its propagation does not occur linearly in the the products diffuse away from the wall to bulk flow. The reactive sys­
porous medium, as the path depends on the preferential flow direction, tems used in their work include aminopolycarboxylic acid group:

7
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

ethylenediaminetetraacetic acid (EDTA), 1,2-cyclohexanediaminetetra­ formation could be described as a function of the inverse Damköhler and
acetic acid (CDTA), and diethylenetriaminepentaacetic acid (DTPA); Péclet number for a randomly structured, highly porous media.
acetic acid (HAc), and HCl with concentrations of 0.25 M and 0.50 M,
respectively. The rocks used in the coreflooding experiments included
7.1. Heterogeneity and porosity
limestone cores measuring 3.8 cm in diameter and 10.2 cm in length,
with porosity between 15 and 20% wt., and permeability ranging from
In general, carbonate reservoirs are heterogeneous at multiple
0.8 to 2mD (Fredd and Fogler, 1998).
dimension scales. Depending on their magnitude, these heterogeneities
Fig. 11 is a classical illustration of reactive flow in limestones re­
could strongly influence the reactive flow in porous media. Here, our
ported in the literature (Fredd and Fogler, 1998). It shows the difference
focus is on discussing the process affecting the core scale, which is the
between the dissolution structures formed after flowing 0.5 M HCl with
domain of most experiments performed to obtain the PVbt and worm­
their respective flow rate and the Damköhler number associated. First,
hole dissolution patterns. In this context, the main heterogeneities of
we can observe that for the same systems, increasing the flow rate will
interest are pore structure, the presence of vugs and microfractures in
decrease the Damköhler number, in the figure reported as NDAmt .
the samples, porosity, and permeability.
Authors observe that plotting the PVbt as a function of the Dam­
In the study developed by (Ziauddin and Bize, 2007), the effect of
köhler number can take into account the reactive contribution of the
pore-scale heterogeneities in carbonate stimulation treatments was
rock-fluid interaction, as shown in Fig. 12.
investigated. Samples of 1.5 inches in diameter and 6 inches in length
Therefore, it is indicative that it is possible to normalize different
cores were used for the experiments. In their study, eight different car­
systems through Damköhler number and estimate the PVbt changing the
bonate rocks were evaluated and classified for Reservoir Rock Type
specific reaction rate. Moreover, it was what they found for a wide range
(RRT). The authors confirm that the response of the carbonate rock to
of the system studied. This could be visualized in Fig. 13, where
acid depends on the RRT to which it belongs. Furthermore, their results
normalized PVbt experimental data were plotted as a function of the
indicate that samples with similar spatial porosity distribution present
inverse of the generalized Damköhler number for a wide range of fluid/
similar PVbt trends.
mineral systems, obtaining an optimal Damköhler.
In the work of Izgec et al. (2009), the acidizing of carbonates of the
They found that this number is approximately equal to 0.29 for a
vuggy structure was studied in 16 carbonate samples, classified as being
wide range of rock-fluid systems. It was then proposed that the optimal
highly heterogeneous with the presence of vugs. Therefore, linear cor­
acid injection rate could be calculated based on this Damköhler number,
eflooding experiments were conducted using plugs with 4 inches in
knowing that both are inversely proportional. In the optimal Damköhler
diameter and 20 inches in length 15% wt. HCl and at room temperature.
number condition, convection, diffusion, and reaction rates are perfectly
We observe that the samples used by authors are bigger than generally
balanced. Only a thin wormhole is formed when the acid is transported
reported in the literature. In this perspective, larger samples reduce the
by convection into the rock.
geometrical effects in the dissolution pattern formed.
Through numerical simulations, Szymczak and Ladd (2009) inves­
The results showed that local pressure drops created by vugs are
tigate the dissolution process in fractures. Authors evaluate the effects of
more dominant in determining the wormhole flow path than chemical
flow rate, reaction rate, and geometrical properties of the fracture to
reactions at the pore level. Furthermore, the acid propagates the
determine the optimal conditions for wormhole formation.
wormholes inside the carbonates of vugular porosity much faster than in
Darcy-scale and convection-diffusion models were used for the re­
more homogeneous rocks. So then, how higher the vuggy concentration
action front and a model for mass transport in the wormhole. The au­
in the sample, the lower the PVbt.
thors observe a relation between the Peclet Number and the inverse of
A critical contribution to the design of acid experiments is a partic­
the Damköhler number, as shown in Fig. 14. Results were summarized in
ular classification of porosity presented in the works of Zakaria et al.
a phase diagram separating different dissolution pattern regimes.
(2015a, 2015b), introducing the concept of the flowing-fraction
Al-Arji et al. (2021), in their work, used diluted hydrochloric acid
parameter (f). It consists of accessible porosity quantification. It is
injected into carbonate rocks (Mount Gambier limestone) to evaluate
defined as the pore volume of injected fluid at which half of the injected
the effect of acid concentrations and injection rates on the wormhole
tracer concentration is observed in the coreflooding outlet flow line.
formation. The dissolution pattern identified were plotted as a function
Fig. 16 shows the plots of dimensionless effluent concentration as a
of dimensionless Damköhler and Péclet numbers, as shown in Fig. 15.
function of the pore volume injected.
Their result confirmed that the critical point for dissolution pattern
In their study, Zakaria et al. (2015a, 2015b) showed that pore

Fig. 11. Different dissolution patterns obtained at different acid flow rates (Fredd and Fogler, 1998).

8
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

Fig. 12. PVbt experimental data of EDTA and HCl plotted as a function of (a) injection rate and (b) inverse of Damköhler number (Fredd and Fogler, 1998).

because it allows relating the pore-scale heterogeneities with the PVbt.


Moreover, the flowing fraction is a non-destructive core flooding test,
allowing the characterization before the reactive flow experiment. For a
given acid, it is possible to correlate the PVbt to another rock type by the
relation presented in Eq. (13).
PVBt1 PVBT2
= Eq. 13
f1 f2
Some values of the flowing fraction are presented in Table 1. We can
observe the value as a reference to the Indiana limestone, the higher
value obtained in the study gives a normalized flowing fraction value of
1.
Authors also test the flowing fraction correlation for vuggy and ho­
mogeneous dolomites with emulsified HCl 15% wt. and demonstrate
that the approach was useful (Zakaria et al., 2015b).
Etten (2015) examined the combined effect of permeability and pore
structure on the acidification efficiency of carbonate matrices. The au­
thors observed that the greater the permeability, the greater would be
the optimal PVbt. Furthermore, when the permeability of the rock rea­
ches a certain value, the influence is reduced. An important point is that
Fig. 13. Normalized PVbt experimental data vs. inverse of Damköhler number the relationship between the two is not linear. Furthermore, they re­
for a wide range of fluid/mineral systems (Fredd and Fogler, 1998). ported that the optimal interstitial velocity is more controlled by pore
structure than permeability only. For similar permeabilities and poros­
ities, tiny, dispersed pores distribution results in a higher optimal
interstitial velocity. Thus, the authors concluded that pore structure is
critical in carbonate acidification design.
Dong et al. (2016) developed a new model based on a statistical
analysis of pore size distribution to calculate optimal conditions in
carbonate acidification. Optimal acid injection rates are calculated
based on semi-empirical flow correlations for different types of flow
geometries. Thus, from the knowledge of the pore size distribution, it is
possible to predict the growth behavior of wormholes. The authors also
show that it is possible to upscale these results for acidification experi­
ments and field treatments.
Alarji et al. (2022) extend the work of Al-Arji et al. (2021) comparing
hard rock limestone and weakly consolidated limestone to validate the
dimensionless approach through Damköhler and Péclet number. One of
their main results is summarized in Fig. 17.
The authors conclude that effective tortuosity has a significant
impact on dissolution patterns. Therefore, recommend extending the
domain diagram, including a third axis representing the effective tor­
Fig. 14. Phase diagram describing the dissolution pattern as a function of
tuosity to capture the heterogeneity of different rock types.
Peclet number and inverse of Damköhler number.

8. Core dimensions
heterogeneity significantly impacts acid response during stimulation
treatment. Their study proved that carbonate rocks with a higher
At the first PVbt experiments performed, there was not much concern
flowing fraction had a higher PVbt than those with a lower flowing
about the dimensions of the sample cores and their influence on the
fraction. The flowing fraction is an essential and usual parameter
optimal PVbt value. However, despite being essential, it is often a

9
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

Fig. 15. Phase diagram obtained in the reactive flow of HCl in Mount Gambier limestone samples (Al-Arji et al., 2021).

morphology of the wormhole, such as optimal injection rate and mini­


mal PVbt. The rocks acidified by 0.7–17% wt. HCl included two types of
rocks: Lavoux and Estaillades, both limestones. Their results showed
that the values of optimal interstitial fluid velocity increased with
increasing core length, as presented in Fig. 18.
Reflecting on the effect observed by Bazin (2001), we noted that the
higher the core length, the higher the time of residence of the acid until
the tip of the wormhole, favoring the contact of the acid with the
wormhole walls. Therefore, more acid is consumed for the same su­
perficial velocity to achieve the breakthrough. Therefore, the optimal
PVbt is expected to present higher values for a longer core sample.
Indeed, the author concluded that the acid volume necessary for
wormhole expansion is given as an order of magnitude. When
comparing wormhole propagation rates at the optimum flow rate, re­
sults reveal that a four-fold increase in wormhole length necessitates a
seven-fold increase in acid volume and injection rate.
Cohen et al. (2008) analyzed the effects of core dimensions and the
optimal conditions for creating wormholes. Their verification found a
contrast in the “confined” and “unconfined” core regimes, defined as
Fig. 16. Flowing fraction plots – tracer concentrations profiles. small (similar to laboratory conditions) and large (similar to field con­
ditions), respectively. The authors’ principal finding is that core
boundary can significantly disrupt the wormholing phenomenon, pri­
Table 1 marily by blocking the wormhole competition mechanism. As a result,
Flowing Fraction values for different carbonates rocks (Theresa acid injection optimization must be carefully handled. The authors point
Sibarani et al., 2018).
out that competition between dissolution rate and convection influences
Rock Type Flowing Fraction the length and diameter of the wormhole created. This competition will
Indiana Limestone 1.00 control the acid distribution along the sidewall and tip. The lowest point
Austin chalk 0.89 of the curve gives the optimal flow conditions for laboratory experi­
Edwards yellow 0.86 ments on the graph of PVbt (Pore Volume to Breakthrough) versus acid
Pink desert 0.73
Winterset limestone 0.61
injection rate. However, what is little diversified is that the core di­
Edwards white 0.53 mensions efficiently influence these optimal conditions. Unconfined
conditions require a lower PVbt value than confined cores, with a dif­
ference of about 130% from the optimal injection rate for the two sit­
neglected feature in experiments. uations (Cohen et al., 2008).
In his study, Buijse (2000) verified the influence of cores geometry Using a two-scale continuum model Kalia and Balakotaiah (2009)
and dimension on wormhole growth. Although there was no quantifi­ studied the impact of medium heterogeneities on the dissolution
cation of this influence, his finding encouraged other researchers to pattern. In the article, initial porosity or permeability fields illustrate
quantitatively detail how much the dimensions of rock cores interfere medium heterogeneity by providing a randomly generated normal dis­
during axial rock experiments. tribution of local porosity values. One exciting element discussed by the
The analysis performed by Bazin (2001) determined that the di­ authors is the effect of aspect ratio. The authors varied the parameter
mensions of the core influence the parameters that govern the

10
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

Fig. 17. Phase diagram comparing the experiments obtained in Al-Arji et al. (2021) and Alarji et al. (2022).

In their work, Dong et al. (2014) studied the effect of the core di­
mensions on the optimum reactive flow conditions in carbonates. Fig. 21
reports their optimal interstitial velocity results as a function of the core
length at different diameters. We observed that for 1 in diameter was not
possible to establish a correlation between the data. However, for the
samples 1.5 and 4 inches is possible to observe that the optimal inter­
stitial velocity does not change above 6 inches.
Between 2 and 6 inches long, the optimal velocity increases with the
length of the cores. Therefore, the authors conclude that as the core
length reaches 6 inches, the vi-opt remains constant and is no longer
affected by the core length.
Dong et al. (2014) found a linear decrease in the volume of acid
needed for the breakthrough with increasing diameter, as shown in
Fig. 22. This result is consistent with previous discussions in the litera­
ture (Furui et al., 2010; K. Furui et al., 2012).
If there is only one dominant wormhole, the diameter of channels
created by the acid in the rock due to dissolution is similar for all cores.
Also, the acid volume for the breakthrough is smaller for cores with
larger diameters (Dong et al., 2014). The work of Dong et al. (2014) is
powerfully relevant to defining a minimum length to work in acidizing
experiments. Nevertheless, few experiments are reported. Therefore, it
is a point that can be explored in new works.
Fig. 18. Effect of core length for 7% HCl, 20 ◦ C, Bazin (2001).
9. Core saturation
α0 = HL for the simulation domain, and maintain the heterogeneity pa­
rameters (permeability and porosity). Their results shown in Fig. 19 Most coreflooding experiments for acidizing start with the rock
show that the number of wormholes that are begun increases with initially fully brine or water (distilled or purified) saturated. This sce­
domain height, but only a few dominant wormholes reach the exit as nario could be representative when several completion fluids are used
dissolution progresses. As a result, as the aspect ratio rises, the amount before acidizing. However, several other scenarios are possible. Some of
of carbonate dissolved as a fraction of the total solid volume falls. them are more representative of the conditions in the reservoir. For
Therefore, the aspect ratio increasing (specifically increasing H with example, the drilling fluid (or its filtrate) oil, and gas also can saturate
L constant), which corresponds to an increase in the diameter in a core the near-wellbore reservoir region. Therefore, it is essential to under­
plug experiment, causes the decrease of the PVbt curve and the optimal stand how other saturation types can affect the PVbt plots and, conse­
PVbt. Other works confirm this result which is possible to see in quently, their influence on operations well. Shukla et al. (2006)
experimental plots in Fig. 20 (K. Furui et al., 2012). However, due to the described in detail in their article some of the saturation conditions
differences in core sample lengths, this result could not be conclusive found in the reservoir near the wellbore region.
regarding the effect of diameter and length. His studies brought an interesting behavior, shown in Fig. 23,

11
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

Fig. 19. Results of dissolution patterns for (a) Da = 5000, (b) Da = 500, (c) Da = 1, (Kalia and Balakotaiah, 2009).

Fig. 20. Results of a linear coreflooding experiment on high-porosity outcrop


chalk samples (K. Furui et al., 2012). Fig. 21. The optimal interstitial velocity under different core lengths for a fixed
core diameter (Dong et al., 2014).
regarding the increment of a gaseous phase before the acid injection.
This resulted in a PVbt reduction of up to 3 times concerning experi­ saturation in carbonate cores, where oils of higher ◦ API facilitate the
ments without gas saturation. Oil saturation also significantly reduces reaction of an acid (HCl) with the pore walls, while oils, such as tar,
the PVbt. The presence of the immiscible phase seems to reduce the acid require a greater volume of acid. Furthermore, when using emulsified
leakage, focusing its reactivity on the wormhole tip. Consequently, a acids in oil-saturated cores, a greater volume of acid is needed for the
lower acid volume is necessary to breakthrough the core sample (Shukla breakthrough due to the reduction of emulsion breakage due to oil
et al., 2006). saturation (Al-Mutairi et al., 2012).
Al-Mutairi et al. (2012) studied a specific scenario of oil saturation in Kumar et al. (2014) studied the impact of residual oil saturation.
tar formations. Some of their results demonstrate that residual oil Their results consolidate previous works, reporting that the residual oil

12
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

the critical point is that the optimal flow rate for full oil saturation is less
than for water, indicating that this saturation needs to be considered in
the treatment design.
All works reported here are important in better understanding the
impact of initial rock saturation. However, it should be noted that the
experiments presented are few and restricted to specific scenarios. Thus,
it is difficult to interconnect them, as well as to perform conclusive
analyses. Therefore, we believe more experiments involving the initial
saturation condition are essential.

10. Temperature

Usually, the temperature is linearly dependent on depth in a


Fig. 22. Effect of the core diameter on the ideal PVbt-opt (Dong et al., 2014). geological formation, increasing with vertical depth. Therefore, chal­
lenging scenarios are found in a geological formation that presents
abnormal temperature and pressure behavior as high-pressure-high-tem­
perature (HTHP) wells. All these temperature variations need to be
counted in well operations once different scenarios can be found in the
porous media.
In the carbonate acidizing process, the temperature mainly affects
the reaction kinetics and diffusion coefficient. It is important to
remember that most chemical reactions are affected by temperature
once it is related to molecules’ agitation degree. Increasing temperature
favors more collisions and, consequently, increases the reactive process.
We want to understand how PVbt will change with temperature in this
context. Fredd and Fogler (1999) experimentally evaluated the effect of
temperature on PVbt plots of HCl 0.5 M. Their results are shown in
Fig. 25.
The results in Fig. 25 show that the higher the temperature, the
higher the optimal PVbt, and the optimal flow rate is shifted to the right.
In other words, we need a higher volume of acid and a higher flow rate
to achieve the optimum point for higher temperatures. The same
behavior was presented by the PVbt curves of HCl 15% wt. at different
temperatures (Fig. 26) obtained through numerical simulations per­
Fig. 23. vol of acid to break the wormhole: only with water saturation, with formed with our in-house software based on Ali and Ziauddin (2020).
water and gas saturation, and with water and oil saturation before acid injec­ Again, the PVbt curve is shifted up and the optimal point to the right.
tion (Shukla et al., 2006). Raising the injected acid’s temperature decreased the wormhole’s
radius in limestone but increased its effectiveness in dolomite. However,
saturation reduces the optimal PVbt compared to the thoroughly at low acid temperatures during dolomite, the acid could not form
water-saturated cores. One of their results is shown in Fig. 24. effective wormholes (Aljawad et al., 2021).
An important point is that there was no observed face dissolution at Investigations by Dong et al. (2018) showed the effects of tempera­
low flow rates. With the increase of the injection rate in the cores ture on different formations. The authors observed an increase in the
saturated with residual oil, the PVbt decreased and showed more minor ideal injection rate for the formation of the dominant wormhole as the
ramifications. One highlight of this work is that an optimal point was not temperature increased. In addition, PVbt slight increase in limestones,
thoroughly characterized. The curve obtained for initial complete oil while there was the opposite in dolomites. Fig. 27 shows the PVbt plots
saturation shows a reduction of PVbt at low flow rates, but the optimal for different temperatures for Silurian Dolomite rock samples. Again, we
point remains above the fully saturated initial condition. Nevertheless, see a completely different pattern compared with HCl.
For mixed formations (limestone and dolomite), there is an increase

Fig. 24. Comparison of acid pore volume with disruption in water-saturated


cores, oil-saturated cores, and residual oil cores from flood water (Kumar Fig. 25. Experimental results of temperature influence on breakthrough curves
et al., 2014). with 0.5 M HCl (Fredd and Fogler, 1999).

13
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

et al., 2017; Xue et al., 2018), as we can observe in Fig. 28.


Aljawad et al. (2021) obtained similar results to those previously
obtained by Li et al. (2017), including the effect on the dolomite for­
mations. Again, as shown in Fig. 29, we see that the heat of the reaction
caused the efficiency curve to shift down for dolomite formations. It is
important to note that the slow reaction rate of dolomite made it more
difficult for acid to breakthrough.
Another important aspect regarding temperature is that the corro­
sion reaction is still favored. Therefore, environments with high tem­
peratures are highly corrosive.
Understanding the effect of temperature helps to design optimized
fluid formulation. Moreover, avoid failures in treatments based on not
appropriately settled experiments.

11. Other variables that may affect carbonate acidizing

Due to the high complexity of carbonate acidizing and the difficulty


of addressing some behaviors, other parameters have been receiving
attention in research. In this section, we report some examples of studies
about specific factors that can affect the reactive flow.
A corrosion Inhibitor is a chemical that slows the acid attack on a
Fig. 26. Simulation results for HCl 15% wt. at different temperatures.
metal surface. It is extensively used in acid formulations. Typically,
cationic organic compounds form a thin physical barrier by adsorption,
acting as retarder additives (Dong et al., 2014; Rabie and Nasr-El-Din,
2015). This behavior affects the PVbt values, as can be seen by
comparing data from Dong et al. (2014) without corrosion inhibitor and
Zakaria et al. (2015a) with 1 vol% with other variables (rock, core di­
mensions, and acid concentration) kept constants. In addition to
corrosion inhibitors, several chemical additives, especially surfactants,
are extensively used in fluid formulations for well-stimulation and
reservoir-enhanced oil recovery (Yarveicy and Haghtalab, 2017; Yar­
veicy and Javaheri, 2019). Thus, a careful analysis must be performed
for each fluid formulation used.
Another aspect is that carbonate dioxide is one of the products during
the classical carbonate-acid reaction. Since this is a gaseous phase in
typical pressures and temperatures, its solubility in spent acid should be
considered. It is common for acidizing studies to be carried out with at
least 1100 psi as backpressure to keep the CO2 solubilized even though
the solubility is very limited, especially in the presence of dissolved
solids. Qiu et al. (2013) concluded in their investigation that, at the same
condition, the presence of gaseous CO2 would be negative to wormhole
formation, considering its interference on reaction rate. Cheng et al.
Fig. 27. Wormhole efficiency ratios in dolomite at 122◦ F, 185◦ F, and 260◦ F (2017) developed a study about the influence of backpressure from 500
(Dong et al., 2018). to 3000 psi on PVbt and wormhole patterns using HCl 15% wt. in

in PVbt (below 50 ◦ C). In these cases, acid is diverted to limestone for­


mations, inhibiting the acid reaction with dolomite (low and moderate
temperatures). However, at high temperatures, the acid invades both
formation types equally. It was also observed that the wormhole extends
for mixed formations even at low temperatures (compared to pure
dolomite). There is a similarity in channel characteristics for high tem­
peratures with pure calcite formations (Aljawad et al., 2021).
At this point of discussion, it is necessary to recall that the wormhole
formation during the reactive flow is a balance between macroscopic
transport (advection), diffusion, and reaction. Therefore, a temperature
change will shift the balance of these contributions and consequently
change the PVbt and the knowledge of the ideal conditions at which
wormhole propagation in a defined temperature is essential for field
applications (Aljawad et al., 2021).
Another aspect related to the temperature is that the reaction
carbonate-HCl releases heat (exothermic reaction). According to nu­
merical experiments reported by Aljawad et al. (2021) there was an
increase of 5 ◦ C in the zones surrounding the wormhole, assuming a
radial flow. However, it appears as though the heat of the reaction has a
minor influence on wormhole propagation (Aljawad et al., 2021; Li Fig. 28. Effects of reaction heat on breakthrough curves at 300 K. (Li
et al., 2017).

14
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

Fig. 29. Simulated acid efficiency curves in limestone (left) and dolomite (right), highlighting the role of reaction heat. (Aljawad et al., 2021).

Indiana Limestone cores. They found that CO2 influence on wormhole Credit author statement
production depends on temperature, pressure, and injection flow rate.
Low flow rates than optimum PVbt conditions are strongly affected by C. R. S. Lucas: Conceptualization, Methodology, Supervision,
backpressure from 75 ◦ F to 150 ◦ F, while in higher flow rates, this effect Writing – original draft. J. R. Neyra.: Writing – original draft E. A.
is less pronounced. Araújo: Writing – original draft. D. N. N. da Silva: Writing – original
Confining pressure represents the overburden pressure e should be draft. M. A. Lima: Writing – original draft. D. A. M. Ribeiro: Writing –
kept constant during the experiment. According to Dong et al. (2014), original draft. Supervision, Writing – review & editing. P.T.P. Aum:
confining pressure is applied to rock during acidizing coreflooding tests Conceptualization, Methodology, Supervision, Writing – original draft,
to avoid the fluid bypassing the core, and usually, it is kept 400 psi Writing – review & editing.
higher than injection pressure. Confining pressure represents the over­
burden pressure e should be kept constant during the experiment. Au­
thors usually use values low as 1200 psi to 5000 psi. The effect of this Declaration of competing interest
parameter on PVbt is not well explored according to our literature
review. The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
12. Conclusions the work reported in this paper.

Knowing the parameters that affect the PVbt is essential in designing Data availability
matrix acidizing operations. In this review article, we started discussing
the theory of reactive flow in porous media focused on carbonated No data was used for the research described in the article.
acidizing and wormholing formation. Then we discuss how to obtain the
PVbt plots experimentally. After, we pass to the main part of the article, Acknowledgments
where we discuss the factors that affect wormhole formation, focusing
on the PVbt plots. Finally, our main conclusions are summarized below. The authors thank the Science and Petroleum Engineering Lab
(LCpetro/UFPA).
• The increase in acid concentration moves down the PVbt plot,
reducing the volume to the breakthrough and shifting the optimal References
flow rate.
• The flow rate is deeply related to the type of wormhole formed. We Abdrazakov, D., Panga, M.K.R., Daeffler, C., Tulebayev, D., 2018. New single-phase
retarded acid system boosts production after acid fracturing in Kazakhstan. In:
discuss the different patterns of dissolution that can be obtained: face
Proceedings - SPE International Symposium on Formation Damage Control. https://
or compact dissolution, conical, wormhole, ramified, and uniform. doi.org/10.2118/189559-ms.
• The reaction rate places a fundamental role in wormhole formation. Adenuga, O.O., Nasr-El-Din, H.A., Sayed, M.A.I., 2013. Reactions of simple organic acids
The dimensionless approach by Damköhler and Peclet numbers and chelating agents with dolomite. In: SPE Production and Operations Symposium.
Society of Petroleum Engineers.
could englobe advection, diffusive and reactive contribution. Ahoulou, A., Tinet, A.-J., Oltéan, C., Golfier, F., Ahoulou, A.W.A., 2020. Experimental
• The geometry in which the reactive flow happens completely insights into the interplay between buoyancy, convection and dissolution reaction
changes the dissolution pattern. In addition, we observe that the core experimental insights into the interplay between buoyancy, convection, and
dissolution reaction. J. Geophys. Res. Solid Earth 1125, 10. https://doi.org/
diameter and length in the experimental approach affect the PVbt. 10.1029/2020JB020854ï.
However, regarding the few experiments reported in the literature Akanni, O.O., Nasr-El-Din, H.A., 2016. Modeling of wormhole propagation during matrix
above 6 inches, the core length will not affect the PVbt value. acidizing of carbonate reservoirs by organic acids and chelating agents. In:
Proceedings - SPE Annual Technical Conference and Exhibition. https://doi.org/
• Core saturation will affect the PVbt plots differently depending on 10.2118/181348-ms.
the type of saturation (residual, partial, or fully saturated) and fluid. al Moajil, A., Caliskan, S., Al-Salem, A., Al-Yami, I., 2019. Aqueous alternative system to
• The temperature has different behavior for limestones and straight and emulsified HCl acids for carbonate acidizing. In: Proceedings - SPE
International Symposium on Oilfield Chemistry. https://doi.org/10.2118/193551-
dolomites. ms.
al Moajil, A., Aljuryyed, N., Alghamdi, S., Caliskan, S., 2020. Performance comparison of
This review is not conclusive once we have very few experiments on retarded acid with emulsified and HCl/formic acid recipes for carbonate acidizing.
In: Society of Petroleum Engineers - Abu Dhabi International Petroleum Exhibition
some of the effects reported in the literature. In addition, we could have
and Conference 2020. https://doi.org/10.2118/203122-MS. ADIP 2020.
other specific factors that can affect the wormhole formation process. Al-Arji, H., Al-Azman, A., Le-Hussain, F., Regenauer-Lieb, K., 2021. Acid stimulation in
Nevertheless, we believe this review could help better understand the carbonates: a laboratory test of a wormhole model based on Damköhler and Péclet
acidizing process, focusing on the PVbt plots obtained in the laboratory. numbers. J. Pet. Sci. Eng. 203, 108593 https://doi.org/10.1016/j.
petrol.2021.108593.

15
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

Al-Mutairi, S. H., Al-Obied, M. A., Al-Yami, I. S., Shebatalhamd, A. M., & Al-Shehri, D. A., Andreasson, E M, Egeli, F, Holmberg, K A, Nystroem, B, Stridh, K G, and Oesterberg, E M.
2012. Wormhole Propagation in Tar During Matrix Acidizing of Carbonate 1993. "Acidizing method using microemulsion. Fremgangsmaate ved syrebehandling
Formation. In SPE international symposium and exhibition on formation damage control. av en underjordisk formasjon." Norway.
OnePetro. Etten, Jordan Ruby (2015). Experimental Investigation on the Effect of Permeability on
Alarji, H., Alazman, A., Regenauer-Lieb, K., 2022. The impact of effective tortuosity on the Optimum Acid Flux in Carbonate Matrix Acidizing. Master’s thesis, Texas A & M
carbonate acidizing and the validation of Damköhler and Péclet dimensionless phase University. Available electronically from https : / /hdl .handle .net /1969 .1
space. J. Pet. Sci. Eng. 212, 110313 https://doi.org/10.1016/j.petrol.2022.110313. /155227.
Ali, M., Ziauddin, M., 2020. Carbonate acidizing: a mechanistic model for wormhole Fredd, C.N., Fogler, H.S., 1996. Alternative stimulation fluids and their impact on
growth in linear and radial flow. J. Pet. Sci. Eng., 106776 https://doi.org/10.1016/j. carbonate acidizing. In: SPE Formation Damage Control Symposium. Society of
petrol.2019.106776. Petroleum Engineers.
Ali, S.A., Ermel, E., Clarke, J., Fuller, M.J., Xiao, Z., Malone, B.P., 2008. Stimulation of Fredd, C.N., Fogler, H.S., 1997. Chelating agents as effective matrix stimulation fluids for
high-temperature sandstone formations from West Africa with chelating agent-based carbonate formations. In: International Symposium on Oilfield Chemistry. Society of
fluids. SPE Prod. Oper. 23, 32–38. https://doi.org/10.2118/93805-pa. Petroleum Engineers.
Ali, S.A., Kalfayan, L., Montgomery, C., 2016. Acid Stimulation. Society of Petroleum Fredd, C.N., Fogler, H.S., 1998. Influence of transport and reaction on wormhole
Engineers, Richardson, Texas, USA. formation in porous media. AIChE J. 44, 1933–1949. https://doi.org/10.1002/
Aljawad, M.S., Aboluhom, H., Schwalbert, M.P., Al-Mubarak, A., Alafnan, S., aic.690440902.
Mahmoud, M., 2021. Temperature impact on linear and radial wormhole Fredd, C., Fogler, H., 1999. Optimum Conditions for Wormhole Formation in Carbonate
propagation in limestone, dolomite, and mixed mineralogy. J. Nat. Gas Sci. Eng. 93 Porous Media: Influence of Transport and Reaction.
https://doi.org/10.1016/j.jngse.2021.104031. Fredd, C.N., Hoefner, M.L., Fogler, H.S., 2017. Microemulsion applications in carbonate
Arruda, G.M., da Silva, D.C., de Azevedo, G.S., Galvão, E.R.V.P., Rodrigues, M.A.F., reservoir stimu-lation. Properties and Uses of Microemulsions 31.
Wanderley Neto, A. de O., 2022. Physicochemical evaluation of the use of alcoholic Furui, K., Burton, R., Burkhead, D., Abdelmalek, N., Hill, A., Zhu, D., Nozaki, M., 2010.
micellar solutions containing nonylphenol and ethanol for the acidizing of carbonate A Comprehensive Model of High-Rate Matrix Acid Stimulation for Long Horizontal
matrices. Colloids Surf. A Physicochem. Eng. Asp. 652 https://doi.org/10.1016/j. Wells in Carbonate Reservoirs.
colsurfa.2022.129821. Furui, K., Burton, R.C., Burkhead, D.W., Abdelmalek, N.A., Hill, A.D., Zhu, D.,
Aum, P.T.P., Souza, T.N., Aum, Y.K.P.G., Dantas, T.N. de C., Neto, A.A.D., 2016. New Nozaki, M., 2012. A Comprehensive Model of High-Rate Matrix-Acid Stimulation for
acid O/W microemulsion systems for application in carbonate acidizing. Int. J. Adv. Long Horizontal Wells in Carbonate Reservoirs Part I-Scaling up Core-Level Acid
Sci. Tech. Res. Iss. 6. Wormholing to Field Treatments.
Aum, P.T.P., Gurgel Aum, Y.K.P., de Andrade Araújo, E., de Almeida Cavalcante, L., Ge, J., Sun, X., Liu, R., Wang, Z., Wang, L., 2020. Emulsion acid diversion agents for oil
Nobre Nunes da Silva, D., Regis dos Santos Lucas, C., de Castro Dantas, T.N., 2021. wells containing bottom water with high temperature and high salinity. ACS Omega
Evaluation of oil-in-water microemulsion base ethoxylated surfactant under acid 5, 29609–29617. https://doi.org/10.1021/acsomega.0c04767.
conditions. Fuel 290. https://doi.org/10.1016/j.fuel.2020.120045. Harris, O.E., Hendrickson, A.R., Coulter, A.W., 1966. High-concentration hydrochloric
Bazin, B., 2001. From matrix acidizing to acid fracturing: a laboratory evaluation of acid/ acid aids stimulation results in carbonate formations. J. Petrol. Technol. 18 (10),
rock interactions. SPE Prod. Facil. 16, 22–29. https://doi.org/10.2118/66566-PA. 1291–1296.
Bazin, B., Abdulahad, G., 1999. Experimental Investigation of Some Properties of Hassan, A., Mahmoud, M., Bageri, B.S., Aljawad, M.S., Kamal, M.S., Barri, A.A.,
Emulsified Acid Systems for Stimulation of Carbonate Formations. https://doi.org/ Hussein, I.A., 2020. Applications of chelating agents in the upstream oil and gas
10.2118/53237-ms. industry: a review. Energy Fuel. 34, 15593–15613. https://doi.org/10.1021/acs.
Bernadiner, M.G., Thompson, K.E., Fogler, H.S., 1991. Effect of Foams Used during energyfuels.0c03279.
Carbonate Acidizing. https://doi.org/10.2118/21035-pa. Hoefner, M.L., Fogler, H.S., 1985. Effective matrix acidizing in carbonates using
Beuterbaugh, A.M., Reyes, E.A., Smith, A.L., 2015. Tandem acidizing-corrosion microemulsions. Chem. Eng. Prog. 81 (United States).
inhibition with low risk-low toxicity chelant. Proc. - SPE Int. Sysposium Oilfield Hoefner, M.L., Fogler, H.S., 1987. Role of acid diffusion in matrix acidizing of
Chem. 1, 440–449. https://doi.org/10.2118/173740-ms. carbonates. J. Petrol. Technol. 39, 203–208. https://doi.org/10.2118/13564-PA.
Buijse, M.A., 2000. Understanding wormholing mechanisms can improve acid treatments Hoefner, M.L., Fogler, H.S., 1988. Pore evolution and channel formation during flow and
in carbonate formations. SPE Production & Facilities 15 (03), 168–175. reaction in porous media. AIChE J. 34 (1), 45–54.
Buijse, M., de Boer, P., Breukel, B., Klos, M., Burgos, G., 2003. Organic acids in carbonate Hongping, Q., Xuele, Z., Qiangying, L., Linyuan, W., Silong, J., 2022. The effect of
acidizing. In: SPE European Formation Damage Conference. Society of Petroleum hydrophobic chains on retarding performance of retarding acids. RSC Adv. 12
Engineers. https://doi.org/10.1039/d2ra00552b.
Carvalho, R.T.R., Oliveira, P.F., Palermo, L.C.M., Ferreira, A.A.G., Mansur, C.R.E., 2019. Huang, T., Hill, A.D., Schechter, R.S., 2000. Reaction rate and fluid loss: The keys to
Prospective acid microemulsions development for matrix acidizing petroleum wormhole initiation and propagation in carbonate acidizing. Spe J. 5 (03), 287–292.
reservoirs. Fuel 238, 75–85. Huang, T., McElfresh, P.M., Gabrysch, A.D., 2003. Carbonate matrix acidizing fluids at
Chacon, O.G., Pournik, M., 2022. Matrix Acidizing in Carbonate Formations. Processes high temperatures: acetic acid, chelating agents or long-chained carboxylic acids?
10, 174. https://doi.org/10.3390/pr10010174. SPE - European formation damage conference. Proceedings, EFDC. https://doi.org/
Cheng, H., Zhu, D., Hill, A.D., 2017. The effect of evolved CO2 on wormhole propagation 10.2118/82268-ms.
in carbonate acidizing. SPE Prod. Oper. 32, 325–332. https://doi.org/10.2118/ Husen, A., Ali, A., Frenier, W.W., Xiao, Z., Ziauddin, M., 2002. Chelating agent-based
178962-PA. fluids for optimal stimulation of high-temperature wells. In: Proceedings - SPE
Cohen, C.E., Ding, D., Quintard, M., Bazin, B., 2008. From pore scale to wellbore scale: Annual Technical Conference and Exhibition, pp. 345–359. https://doi.org/
impact of geometry on wormhole growth in carbonate acidization. Chem. Eng. Sci. 10.2523/77366-ms.
63, 3088–3099. https://doi.org/10.1016/j.ces.2008.03.021. Izgec, O., Zhu, D., & Hill, A. D. (2009). Models and methods for understanding of early
Crowe, C.W., McGowan, G.R., Baranet, S.E., 1990. Investigation of retarded acids acid breakthrough observed in acid core-floods of vuggy carbonates. In 8th European
provides better understanding of their effectiveness and potential benefits. SPE Prod. Formation Damage Conference. OnePetro.
Eng. 5 https://doi.org/10.2118/18222-pa. Izgec, O., Zhu, D., Hill, A.D., 2010. Numerical and experimental investigation of acid
Daccord, G., Lietard, O., Lenormand, R., 1993. Chemical dissolution of a porous medium wormholing during acidization of vuggy carbonate rocks. J. Pet. Sci. Eng. 74, 51–66.
by a reactive fluid—II. Convection vs reaction, behavior diagram. Chemical https://doi.org/10.1016/j.petrol.2010.08.006.
engineering science 48 (1), 179–186. Kalfayan, L., 2008. Production Enhancement with Acid Stimulation, second ed. Pennwell
Dantas, T.N.C., Santanna, V.C., Souza, T.T.C., Lucas, C.R.S., Dantas Neto, A.A., Aum, P.T. Books, Oklahoma.
P., 2019. Microemulsions and nanoemulsions applied to well stimulation and Kalia, N., Balakotaiah, V., 2009. Effect of medium heterogeneities on reactive dissolution
enhanced oil recovery (EOR). Brazilian J. Petrol. Gas 12. https://doi.org/10.5419/ of carbonates. Chem. Eng. Sci. 64, 376–390. https://doi.org/10.1016/j.
bjpg2018-0023. ces.2008.10.026.
de Antuñano, Y., Losada, M., Milne, A., 2015. Stimulating high-temperature dolomitic Khan, M., Qamruzzaman, M., Roy, D.C., Raman, R., 2021. Mitigation of high
limestone reservoirs with chelant fluids. In: SPE - European Formation Damage temperature challenges in limestone acidizing through the use of chelating agents.
Conference, Proceedings, EFDC 2015-Janua, pp. 1218–1240. https://doi.org/ In: Proceedings - SPE Annual Technical Conference and Exhibition. https://doi.org/
10.2118/174167-ms. 10.2118/206039-MS.
de Castro Dantas, T.N., de Oliveira, A.C., de Souza, T.T.C., dos Santos Lucas, C.R., de Kumar, R., He, J., Texas, A., Members, S.P.E., 2014. Effect of Oil Saturation on Acid
Andrade Araújo, E., Aum, P.T.P., 2019. Experimental study of the effects of acid Propagation during Matrix Acidization of.
microemulsion flooding to enhancement of oil recovery in carbonate reservoirs. Kumar, H., Muhemmed, S., Nasr-El-Din, H., 2021. The role of CO2 in carbonate acidizing
J. Pet. Explor. Prod. Technol. https://doi.org/10.1007/s13202-019-00754-x. at the field scale-A multi-phase perspective. In: SPE Annual Technical Conference
Dong, K., Jin, X., Zhu, D., Hill, A.D., 2014. The Effect of Core Dimensions on the and Exhibition. OnePetro.
Optimum Acid Flux in Carbonate Acidizing. Li, S., Li, Z., Lin, R., 2008. Mathematical models for foam-diverted acidizing and their
Dong, K., Zhu, D., Hill, A.D., 2016. Theoretical and experimental study on optimal applications. Petrol. Sci. 5 https://doi.org/10.1007/s12182-008-0022-4.
injection rates in carbonate acidizing. In: SPE Journal. Society of Petroleum Li, Y., Liao, Y., Zhao, J., Peng, Y., Pu, X., 2017. Simulation and analysis of wormhole
Engineers (SPE), pp. 892–901. https://doi.org/10.2118/178961-pa. formation in carbonate rocks considering heat transmission process. J. Nat. Gas Sci.
Dong, K., Zhu, D., Hill, A.D., 2018. The role of temperature on optimal conditions in Eng. https://doi.org/10.1016/j.jngse.2017.02.048.
dolomite acidizing: an experimental study and its applications. J. Pet. Sci. Eng. 165, Maheshwari, P., Ratnakar, R.R., Kalia, N., Balakotaiah, V., 2013. 3-D simulation and
736–742. analysis of reactive dissolution and wormhole formation in carbonate rocks. Chem.
Economides, M., Nolte, Kenneth, 2000. Reservior Stimulation. Wiley, Sugar Land, third Eng. Sci. 90, 258–274. https://doi.org/10.1016/j.ces.2012.12.032.
ed. Wiley, New York. https://doi.org/10.1017/CBO9781107415324.004. Mahmoud, M., Barri, A., Elkatatny, S., 2017. Mixing chelating agents with seawater for
acid stimulation treatments in carbonate reservoirs. J. Pet. Sci. Eng. 152, 9–20.

16
C.R.S. Lucas et al. Journal of Petroleum Science and Engineering 220 (2023) 111168

Oliveira, H.A., Li, W., Maxey, J.E., 2013. Invert emulsion acid for simultaneous acid and Tariq, Z., Aljawad, M.S., Hassan, A., Mahmoud, M., Al-Ramadhan, A., 2021. Chelating
proppant fracturing. In: Proceedings of the Annual Offshore Technology Conference, agents as acid-fracturing fluids: experimental and modeling studies. Energy Fuel. 35,
1, pp. 522–530. https://doi.org/10.4043/24332-ms. 2602–2618. https://doi.org/10.1021/acs.energyfuels.0c04045.
Paccaloni, G., Tambini, M., Galoppini, M., 1988. Key Factors for Enhanced Results of Theresa Sibarani, T., Ziauddin, M., Nasr-El-Din, H.A., Zakaria, A.S., 2018. The Impact of
Matrix Stimulation Treatments. https://doi.org/10.2118/17154-ms. Pore Structure on Carbonate Stimulation Treatment Using VES-Based HCl. https://
Pandya, N., Wadekar, S., 2013. A novel emulsified acid system for stimulation of very doi.org/10.2118/192066-MS.
high-temperature carbonate reservoirs. In: Society of Petroleum Engineers - Wanderley Neto, da Silva, Arruda, G.M., da Hora, Rodrigues, M.A.F., 2021. Chemical
International Petroleum Technology Conference 2013, IPTC 2013: Challenging study of the application of nonionic surfactants nonylphenol in delaying the
Technology and Economic Limits to Meet the Global Energy Demand. https://doi. acidizing reaction of carbonate matrices. J. Disper. Sci. Technol. 1–8.
org/10.2523/iptc-16452-ms. Wang, Y., Hill, A.D., Schechter, R.S., 1993. The Optimum Injection Rate for Matrix
Parkinson, M., Munk, T., Brookley, J., Caetano, A., Albuquerque, M., Cohen, D., Acidizing of Carbonate Formations.
Reekie, M., 2010. Stimulation of multilayered high-carbonate-content sandstone Xue, H., Huang, Z., Zhao, L., Wang, H., Kang, B., Liu, P., Liu, F., Cheng, Y., Xin, J., 2018.
formations in West Africa using chelant-based fluids and mechanical diversion. In: Influence of acid-rock reaction heat and heat transmission on wormholing in
Proceedings - SPE International Symposium on Formation Damage Control, 2, carbonate rock. J. Nat. Gas Sci. Eng. 50, 189–204. https://doi.org/10.1016/j.
pp. 769–778. https://doi.org/10.2118/128043-ms. jngse.2017.12.008.
Qiu, X., Zhao, W., Chang, F., Dyer, S., 2013. Quantitative modeling of acid wormholing Yarveicy, H., Haghtalab, A., 2017. Effect of amphoteric surfactant on phase behavior of
in carbonates- what are the gaps to bridge. In: All Days. SPE. https://doi.org/ hydrocarbon-electrolyte-water system-an application in enhanced oil recovery.
10.2118/164245-MS. J. Dispersion Sci. Technol. 39, 522–530. https://doi.org/10.1080/
Rabie, A.I., Nasr-El-Din, H.A., 2015. Effect of acid additives on the reaction of 01932691.2017.1332525.
stimulating fluids during acidizing treatments. In: Day 2 Tue, September 15, 2015. Yarveicy, H., Javaheri, A., 2019. Application of Lauryl Betaine in enhanced oil recovery:
SPE. https://doi.org/10.2118/175827-MS. a comparative study in micromodel. Petroleum 5, 123–127. https://doi.org/
Rae, P., di Lullo, G., 2003. Matrix acid stimulation - a review of the state-of-the-art. In: 10.1016/j.petlm.2017.09.004.
SPE European Formation Damage Conference. https://doi.org/10.2118/82260-MS. Yarveicy, H., Habibi, A., Pegov, S., Zolfaghari, A., Dehghanpour, H., 2018. Enhancing oil
Reyes, E.A., Smith, A., Beuterbaugh, A., 2013. Carbonate stimulation with biodegradable recovery by adding surfactants in fracturing water: a montney case study. In: Day 1
chelating agent having broad unique spectrum (pH, Temperature, Concentration) Tue. SPE. https://doi.org/10.2118/189829-MS. March 13, 2018.
activity, 3. In: SPE Middle East Oil and Gas Show and Conference, MEOS, Yoo, H., Nguyen, T., Lee, J., 2022. Improved wormhole prediction model considering
Proceedings, pp. 1758–1768. https://doi.org/10.2118/164380-ms. propagation characteristics of wormhole head in carbonate acidizing. J. Petrol. Sci.
Shafiq, M.U., Mahmud, H. ben, 2017. Sandstone matrix acidizing knowledge and future Eng. 216, 110807.
development. J. Pet. Explor. Prod. Technol. https://doi.org/10.1007/s13202-017- Zakaria, A.S.S., Nasr-El-Din, H.A.A., Ziauddin, M., 2015a. Predicting the performance of
0314-6. the acid-stimulation treatments in carbonate reservoirs with nondestructive tracer
Shukla, S., Zhu, D., Hill, A.D., Texas, U., 2006. The Effect of Phase Saturation Conditions tests. SPE J. 20, 1238–1253. https://doi.org/10.2118/174084-PA.
on Wormhole Propagation in Carbonate Acidizing, pp. 273–281. Zakaria, A.S., Nasr-El-Din, H.A., Ziauddin, M., 2015b. Flow of emulsified acid in
Siddiqui, S., Nasr-El-Din, H.A., Khamees, A.A., 2006. Wormhole initiation and carbonate rocks. Ind. Eng. Chem. Res. 54, 4190–4202. https://doi.org/10.1021/
propagation of emulsified acid in carbonate cores using computerized tomography. ie504167y.
J. Pet. Sci. Eng. 54 https://doi.org/10.1016/j.petrol.2006.08.005. Ziauddin, E.M., Bize, Emmanuel, 2007. The effect of pore scale heterogeneities on
Szymczak, P., Ladd, A.J.C., 2009. Wormhole formation in dissolving fractures. carbonate stimulation treatments. In: SPE Middle East Oil and Gas Show and
J. Geophys. Res. Solid Earth 114. https://doi.org/10.1029/2008JB006122. Conference.

17

You might also like