Sei sulla pagina 1di 28

Fundamentals of
Acid Stimulation
A. Daniel Hill and Robert S. Schechter, University of Texas at Austin

16-1. Introduction or other mechanical aspects of the completion. Two


exceptions to this rule may occur. First, in highly pro-
Matrix stimulation is a technique in which a solvent ductive wells, the productivity improvement of about
is injected into the formation to dissolve some of the 20% that is possible with matrix stimulation of an
materials present and hence recover or increase the undamaged well may be economic. Second, in natu-
permeability in the near-wellbore region. Such treat- rally fractured or highly vugular carbonate reservoirs,
ments are called “matrix” treatments because the sol- live acid may penetrate to a sufficient distance to yield
vent is injected at pressures below the parting pressure a productivity enhancement greater than that normally
of the formation so that fractures are not created. The expected from a true matrix treatment.
objective is to greatly enhance or recover the perme- An ideal matrix treatment restores the permeability
ability near the wellbore, rather than affect a large in the near-wellbore region to a value at least as high
portion of the reservoir. as the original undamaged permeability; it accom-
The most common matrix stimulation treatment plishes this over the entire completed interval and it
is acidizing, in which an acidic solution is injected leaves the formation in the treated region with high
to dissolve minerals in the formation. However, other relative permeability to the oil and/or gas phase.
solvents are also used. The next most common fluids Designing a treatment should strive to achieve this
are organic solvents aimed at dissolving waxes, paraf- ideal at the lowest possible cost, which requires con-
fins, asphaltenes or other organic damaging materials. sideration of the many physical and chemical interac-
Nonacid matrix stimulation is addressed in Chapter 14. tions taking place between the injected fluids and the
This chapter focuses on matrix acidizing; however, the reservoir minerals and fluids. The most important of
reader should keep in mind that many of the theories these phenomena are the following:
and calculation procedures presented here can also be
applied to nonacid solvent treatments. • mass transfer of acid molecules to the mineral sur-
The most common acids are hydrochloric acid face and subsequent reaction at the surface—This
(HCl), used primarily to dissolve carbonate minerals, fundamental process of acidizing is illustrated in
and mixtures of HCl and hydrofluoric acid (HF), used Fig. 16-1. Acid reactions with minerals are termed
to attack silicate minerals such as clays and feldspars. heterogeneous reactions because they occur at a
Other acids, particularly some weak organic acids, boundary between the solid and the liquid rather
are used in special applications, such as high-temper- than in the bulk phases. Before the reaction can
ature wells. Matrix acidizing is a near-wellbore treat- occur, acid must be transported to the mineral sur-
ment, with all the acid reacting within about 1 ft of face by convection or diffusion. The overall reac-
the wellbore in sandstone formations and within a tion rate (i.e., the rate of change of the concentra-
few inches to perhaps as much as 10 ft from the well- tion of one component in the bulk liquid phase)
bore in carbonates. may depend on both the rate of mass transfer and
Matrix acidizing can significantly enhance the pro- the rate of surface reaction. Many times, however,
ductivity of a well when near-wellbore formation one of these processes is much slower than the
damage is present and, conversely, is of limited bene- other and controls the overall rate, in which case
fit in an undamaged well, as shown in Chapter 1. the faster process can be ignored.
Thus, matrix acidizing generally should be applied • changing pore structure—The physical change in
only when a well has a high skin factor that cannot be the pore structure caused by dissolution of some of
attributed to partial penetration, perforation efficiency the minerals by acid is the mechanism by which

Reservoir Stimulation 16-1


• fluid selection—acid type, concentration and volume
• injection schedule—planned rate schedule and
sequence of injected fluids
Bulk solution
• acid coverage and diversion—special steps taken to
improve acid contact with the formation
Acid transport Product transport • real-time monitoring—methods to evaluate the
by convection by convection acidizing process as it occurs
or diffusion or diffusion
• additives—other chemicals included in the acid
Acid
solution to enhance the process or to protect tubular
concentration goods.
near the
Heterogeneous surface Treatment design is considered in detail in other
reaction at Reactive mineral
solid/liquid chapters. This chapter lays the foundation for the
interface design methods aimed at optimizing a matrix acidizing
treatment by reviewing the underlying chemistry and
Figure 16-1. Acid reaction occurring in a system. physics of the acidizing process and introducing the
latest models of the processes involved. First, the inter-
action of acids with reservoir minerals is addressed.
matrix acidizing increases permeability. The manner Then, current models of the matrix acidizing process
in which the pore structure changes is fundamental- in sandstones and in carbonates are presented.
ly different in sandstones and carbonates, which
leads to radically different approaches to modeling
the acidizing process in these two mineralogies. 16-2. Acid-mineral interactions
• precipitation of reaction products—Secondary reac-
tions occur in acidizing, particularly in sandstones, 16-2.1. Acid-mineral reaction stoichiometry
that can result in the precipitation of reaction prod-
The amount of acid required to dissolve a given
ucts from the bulk liquid phase. Obviously, precipi-
amount of mineral is determined by the stoichiometry
tated solids may block pore spaces and work
of the chemical reaction, which describes the number
against the goal of matrix acidizing.
of moles of each species involved in the reaction. For
• acid fluid–reservoir fluid interactions—The acid example, the simple reaction between HCl and calcite
solution injected in matrix acidizing may interact (CaCO3) can be written as
physically and/or chemically with the reservoir flu-
ids as well as with the minerals. These interactions 2HCl + CaCO3 → CaCl2 + CO2 + H2O
can result in changes in wettability, phase saturation which shows that 2 moles of HCl are required to dis-
distribution, precipitation of solids or emulsification. solve 1 mole of CaCO3. The numerals 2 and 1 multi-
• variations in reservoir permeability or the distribu- plying the species HCl and CaCO3 are the stoichio-
tion of damage—A successful acidizing treatment metric coefficients νHCl and νCaCO 3 for HCl and
requires contacting all damaged regions around the CaCO3, respectively.
well with acid. This is usually complicated by vari- When HF reacts with silicate minerals, numerous
ations in the injectivity to acid along the wellbore, secondary reactions may occur that influence the over-
which leads to the use of techniques to affect good all stoichiometry of the reaction. For example, when
acid coverage (acid diversion). HF reacts with quartz (SiO2), the primary reaction is
In considering the many aspects of the matrix 4HF + SiO2 SiF4 + 2H2O
acidizing process, the focus is on the key design vari- producing silicon tetrafluoride (SiF4) and water. The
ables; to be useful, any model of the process must aid stoichiometry of this reaction shows that 4 moles of
in optimizing the design. The primary design consid- HF are required to consume 1 mole of SiO2. However,
erations are the SiF4 produced may also react with HF to form flu-
osilicic acid (H2SiF6) according to

16-2 Fundamentals of Acid Stimulation


SiF4 + 2HF H2SiF6 In modeling the acidizing process, it is therefore
crucial to determine the local solution composition
If this secondary reaction goes to completion, 6 moles
as the process progresses. To satisfy this requirement,
of HF, rather than 4 moles, will be consumed to dis-
Sevougian et al. (1992) developed a geochemical
solve 1 mole of quartz. A complication is that the flu-
model that considers a local partial equilibrium among
osilicates may exist in various forms, so that the total
certain reactions involved in the acidizing process.
amount of HF required to dissolve a given amount of
This is an important modeling capability; for example,
quartz depends on the solution concentration.
Shuchart and Gdanski (1996) observed that the
Typical reactions involved in HF acidizing are sum-
AlF2+/AlF 2+ ratio obeys a pseudoequilibrium relation
marized in Table 16-1. Only the primary reaction prod-
over a range of reaction conditions. This implies local
ucts, AlF2+ and SiF4, are shown although many other
equilibrium among the various aluminum fluoride
products are possible (Labrid, 1975; Hekim and Fogler,
species in the aqueous phase. To utilize the full capabil-
1982; Walsh et al., 1982; Sevougian et al., 1992;
ity of the Sevougian et al. model, the rates of both the
Shuchart and Gdanski, 1996). How many moles of
homogeneous and heterogeneous reactions that are not
mineral are dissolved by a mole of acid is an important
fully at equilibrium are required. Not all the important
consideration in the selection of an acid treatment vol-
reaction rates are known.
ume that may differ substantially from the values
Thus, the chemistry of the reactions of HF with the
shown in Table 16-1. As an example of the uncertainty,
minerals found in sandstone formations is complex
Schechter (1992) suggested that for conservative design
and difficult to model. An additional complexity is the
purposes, about 20 moles of acid are required to dis-
tendency for some of the reaction products to precipi-
solve 1 mole of feldspar rather than the 14 moles
tate as the acid reactions go to completion (see
shown in Table 16-1. Depending on the composition of
Section 16-2.3). For practical purposes, it is conve-
the acid solution, the actual value may be substantially
nient to express the stoichiometry in terms of the
less. This is especially the case if the silicon extracted
approximate “dissolving power,” as introduced by
from the feldspar crystals appears finally as the precipi-
Williams et al. (1979).
tate Si(OH)4 rather than as soluble SiF4 as shown in
The dissolving power expresses the amount of min-
Table 16-1. If the final silicon product is, in fact,
eral that can be consumed by a given amount of acid
Si(OH)4, then only 1 mole of HF would be required
on a mass or volume basis. First, the gravimetric dis-
to dissolve a single mole of albite.
solving power β, which is the mass of mineral con-
sumed by a given mass of acid, is defined as
Table 16-1. Primary chemical reactions in acidizing.
ν mineral MWmineral
β= , (16-1)
HCl ν acid MWacid
Calcite 2HCl + CaCO3 → CaCl2 + CO2 + H2O
where the ν terms are the stoichiometric coefficients
Dolomite 4HCl + CaMg(CO3)2 → CaCl2 + MgCl2
+ 2CO2 + 2H2O and MWmineral and MWacid are the molecular weights of
Siderite 2HCl + FeCO3 → FeCl2 + CO2 + H2O
the mineral and the acid, respectively. Thus, for the
reaction between 100% HCl and CaCO3,
HCl-HF
(1)(100.1)
Quartz 4HF + SiO2 SiF4 (silicon tetrafluoride)
β100 = = 1.37 lbm CaCO3 /lbm HCl, (16-2)
+ 2H2O
(2)(36.5)
SiF4 + 2HF H2SiF6 (fluosilicic acid)
where the subscript 100 denotes 100% HCl. The dis-
Albite solving power of any other concentration of acid is
(sodium feldspar) NaAlSi3O8 + 14HF + 2H+ +
Na + AlF +

β100 times the weight fraction of acid in the acid solu-


2
+ 3SiF4 + 8H2O
Orthoclase tion. For the commonly used 15% HCl, β15 = 0.15 ×
(potassium β100 = 0.21 lbm CaCO3/lbm HCl. The stoichiometric
feldspar) KAlSi3O8 + 14HF + 2H+ K+ + AlF2+
+ 3SiF4 + 8H2O
coefficients for common acidizing reactions are found
from the reaction equations in Table 16-1.
Kaolinite Al4Si4O10(OH)8 + 24HF + 4H+ 4AlF2+
+ 4SiF4 + 18H2O The volumetric dissolving power X, which is simi-
Montmorillonite Al4Si8O20(OH)4 + 40HF + 4H+ 4AlF2+
larly defined as the volume of mineral dissolved by a
+ 8SiF4 + 24H2O given volume of acid, is related to the gravimetric dis-
solving power by

Reservoir Stimulation 16-3


ρacid solution 16-2.2. Acid-mineral reaction kinetics
X =β . (16-3)
ρmineral
The reaction between an acid and a mineral occurs
A 15% HCl solution has a specific gravity of about when acid reaches the surface of the mineral by diffu-
1.07 and CaCO3 has a density ρ of 169 lbm/ft3. For sion or convection from the bulk solution. The overall
the reaction of these species, the volumetric dissolving rate of acid consumption or mineral dissolution depends
power is on two distinct phenomena—the rate of transport of
acid to the mineral surface by diffusion or convection
 lbm 15% HCl 
lbm CaCO3   (
1.07)(62.4) 3 and the actual reaction rate on the mineral surface.
X15 = 0.21 ft 15% HCl  Usually, one of these processes is much slower than
 lbm 15% HCl   lbm 15% CaCO3 
 169 3  the other. In this case, the fast process can be ignored,
 ft 15% CaCO3  because it can be thought of as occurring in an insignif-
ft 3 CaCO3 icant amount of time compared with the slow process.
= 0.082 .
ft 3 15% HCl (16-4) For example, the HCl-CaCO3 reaction rate is
extremely high, so the overall rate of this reaction is
The dissolving powers of various acids with limestone usually controlled by the rate of acid transport to the
and dolomite and of HF with quartz and albite are surface, the slower of the two processes. On the other
given in Tables 16-2 and 16-3, respectively (Schechter, hand, the surface reaction rates for many HF-mineral
1992). Sidebar 16A is an example calculation using reactions are slow compared with the acid transport
dissolving power. rate, and the overall rate of acid consumption or min-
eral dissolution is reaction-rate controlled. The “kinetics”

Table 16-2. Dissolving power of various acids (Schechter, 1992).

X (%)
Formation Acid β100 5 10 15 30

Limestone (CaCO3) Hydrochloric (HCl) 1.37 0.026 0.053 0.082 0.175


Formic (HCOOH) 1.09 0.020 0.041 0.062 0.129
Acetic (CH3COOH) 0.83 0.016 0.031 0.047 0.096

Dolomite (CaMg(CO3)2) Hydrochloric 1.27 0.023 0.046 0.071 0.152


Formic 1.00 0.018 0.036 0.054 0.112
Acetic 0.77 0.014 0.027 0.041 0.083

Notes: ρCaCO 3 = 2.71 g/cm3, ρCaMg(CO 3)2 = 2.87 g/cm3

Table 16-3. Dissolving power of hydrofluoric acid (Schechter, 1992).

Acid concentration (wt%) Quartz (SiO2) Albite (NaAlSi2O8)


β X β X

2 0.015 0.006 0.019 0.008


3 0.023 0.010 0.028 0.011
4 0.030 0.018 0.037 0.015
6 0.045 0.019 0.056 0.023
8 0.060 0.025 0.075 0.030

Notes: β = mass of rock dissolved/mass of acid reacted, X = volume of rock dissolved/volume of acid reacted

16-4 Fundamentals of Acid Stimulation


16A. Calculating minimum acid volume using dissolving power

The volume of acid required for a matrix acidizing treatment can be estimated using the concept of dissolving power. Because the vol-
umetric dissolving power X is the volume of a particular mineral that is dissolved by a given volume of a particular acid solution, the
minimum acid requirement to remove that mineral can be calculated with little information other than the dissolving power. Consider
the following problem:
A sandstone formation with a porosity of 0.2 contains 5-vol% albite (sodium feldspar). What is the minimum volume of 3% HF solu-
tion required to dissolve all the albite a distance of 6 in. beyond a 6-in. diameter wellbore?

Solution
The minimum acid volume is the amount VHF required to dissolve all the feldspar plus the amount Vp required to fill the pore space in
the region of feldspar dissolution. These volumes are

Vfeldspar = π(rHF2 − rw2 )(1− φ)x feldspar = π(0.75 2 − 0.25 2 )(1− 0.2)(0.05)
ft 3 (16A-1)
= 0.063 feldspar
ft

Vfeldspar 0.063 ft 3 HF
VHF = = = 5.7 . (16A-2)
X3 0.011 ft

In these equations, rHF is the radial penetration distance of HF, rw is the wellbore radius, φ is the porosity, and xfeldspar is the volume frac-
tion of the sandstone that is feldspar. The volume of pore space within 6 in. of the wellbore after removal of the feldspar is

( ) (
Vp = π(rHF2 − rw2 ) φ + x feldspar (1− φ) = π(0.75 2 − 0.25 2 ) 0.2 + 0.05(1− 0.2) )
ft 3 (16A-3)
= 0.38 ,
ft
so the total volume of HF required is

ft 3  7.48 gal  gal


VHF ,T = VHF + V p = (5.7 + 0.38)   = 46 . (16A-4)
ft  ft 3  ft

Thus, the minimum volume of 3% HF solution required to remove all feldspar in a radial region extending 6 in. beyond the wellbore
is 46 gal/ft of reservoir thickness. In an actual acidizing treatment, the injected acid does not react with feldspar only, and as shown by
examining models of the acidizing process, the acid is not spent uniformly, as tacitly assumed in this calculation. Nevertheless, this
simple calculation provides a ballpark figure for acid requirements and is a handy check of more complex models of the process.

of a reaction is a description of the rate at which the eral RB, which is related to the acid consumption rate
chemical reaction takes place, once the reacting through the stoichiometry of the reaction
species have been brought into contact. ν
A reaction rate is generally defined as the rate of RA = A RB , (16-6)
νB
appearance in the solution of the species of interest in
units of moles per second (mol/s). A surface reaction where νA and νB are the stoichiometric coefficients for
rate depends on the amount of surface exposed to acid A and mineral B.
reaction, so these reactions are expressed per unit of The reaction rate rA generally depends on the con-
surface area. In general, the surface reaction rate of an centrations of the reacting species. However, in the
aqueous species of acid A reacting with mineral B is reaction between an aqueous species and a solid, the
concentration of the solid can be ignored, because it
RA = rA SB , (16-5) remains essentially constant. For example, a grain of
quartz has a fixed number of moles of quartz per unit
where RA is the rate of appearance of acid A in mol/s,
volume of quartz, irrespective of reactions that may
rA is the surface area-specific reaction rate of A in
be occurring on the surface of the grain. Incorporating
mol/s-m2, and SB is the surface area of mineral B.
concentration dependence into the rate expression
When A is being consumed, the reaction rates rA and
yields
RA are negative. Acid-mineral reaction rates are typi-
cally expressed as the rate of dissolution of the min- − RA = E f CAα SB , (16-7)

Reservoir Stimulation 16-5


where Ef is a reaction rate constant in mol A/[m2-s- The constants α, Efo and ∆E/R are listed in
(mol A/m3)α], CA is the concentration of species A at Table 16-4. SI units are used in these expressions,
the reactive surface, and α is the order of the reaction so CHCl is in kg-mol/m3 and temperature T is in
(i.e., a measure of how strongly the reaction rate degrees Kelvin. The reaction rate rHCl is expressed
depends on the concentration of A). The reaction rate as kg-mol HCl reacted/m2-s.
constant depends on temperature and sometimes on
• Reactions of hydrochloric and weak acids with
the concentration of chemical species other than A.
carbonates
Finally, Eq. 16-7 is written in the conventional manner
for a species of acid that is being consumed from HCl is a strong acid, meaning that when HCl is
solution, by placing a minus sign with RA so that dissolved in water, the acid molecules almost com-
Ef is a positive number. pletely dissociate to form hydrogen ions (H+) and
chloride ions (Cl–). The reaction between HCl and
• Laboratory measurement of reaction kinetics carbonate minerals is actually a reaction of the H+
To measure the surface reaction rate of acid-mineral with the mineral. With weak acids, such as acetic
reactions, it is necessary to maintain a constant or formic acid, the reaction is also between H+ and
mineral surface area or measure its change during the mineral, with the added complication that the
reaction and to ensure that the rate of acid transport acid is not completely dissociated, thus limiting the
to the mineral surface is fast relative to the reaction supply of H+ available for reaction. Because H+ is
rate. The two most common methods of obtaining the reactive species, the kinetics of the HCl reac-
these conditions are with a well-stirred slurry of tion can also be used for weak acids by considering
mineral particles suspended in an acid solution (a the acid dissociation equilibrium.
stirred reactor) or with a rotating disk apparatus The kinetics of a weak acid–carbonate mineral
(Fogler et al., 1976.) In the rotating disk apparatus, reaction may therefore be obtained from Eq. 16-8
a disk of the mineral is placed in a large container as follows (Schechter, 1992):
holding the acid solution. The disk is rotated rapidly,
− rweak acid = E f K dα / 2 Cweak
α/2
acid , (16-10)
so that the acid mass-transfer rate is high relative
to the surface reaction rate. A third, more indirect where Kd is the dissociation constant of the weak
method is by matching the coreflood response to acid and Ef is the reaction rate constant for the HCl-
acidizing with a model of the process of flow with mineral reaction.
reaction.
Lund et al. (1975, 1973) measured the kinetics • Reactions of hydrofluoric acid with sandstone
of the HCl-calcite and HCl-dolomite reactions, minerals
respectively. Their results may be summarized as HF reacts with virtually all of the many mineral
constituents of sandstone. Reaction kinetics have
− rHCl = E f CHCl
α
(16-8)
been reported for the reactions of HF with quartz
(Bergman, 1963; Hill et al., 1977), feldspars
∆E 
E f = E of exp − . (16-9) (Fogler et al., 1975) and clays (Kline and Fogler,
 RT 

Table 16-4. Constants in HCl-mineral reaction kinetics models.

∆E
Mineral α Ef o (K)
R
 
 kg - mol HCl 
 α

 2  kg - mol HCl  
 m -s 3  
 m acid solution  

Calcite (CaCO3) 0.63 7.291 × 107 7.55 × 103

Dolomite (CaMg(CO3)2) 6.18 × 10 −4T 9.4 × 1011 11.32 × 103


1− 2 × 10 −3T 1000 α

16-6 Fundamentals of Acid Stimulation


1981a). These kinetic expressions can all be repre- tact with the acidic solution. Therefore, the mor-
sented by phology of the mineral assemblage becomes

[ ]
an important issue.
− rmineral = E f 1 + K (CHCl ) CHF
β α
, (16-11) Comparison of the reaction rates of various min-
erals requires placing the rates on the basis of a unit
for which the parameters α, β, Ef and the empirical of reactive area. On this basis, montmorillonites
kinetic constant K are listed in Table 16-5. and kaolinites react about 2 orders of magnitude
These expressions show that the dependence on slower than feldspars, and illites react at least 1
HF concentration is approximately first order (α = order of magnitude slower than kaolinite. Viewing
1). For feldspar reactions, the reaction rate increases thin sections of rocks following acid treatment with
with increasing HCl concentration, although HCl is HCl-HF mixtures shows that the feldspars are usu-
not consumed in the reaction. Thus, HCl catalyzes ally removed because of their high specific reaction
HF-feldspar reactions. Kline and Fogler (1981a) rates. Authogenic clays also appear to react rapidly
showed that the reactive area depends on the crystal- because of their intimate exposure to the acidic
line structure of the clay reacting with an HCl-HF solution. On the other hand, clastic clays are com-
mixture and is generally only a small fraction of monly found in thin sections following acid treat-
the total surface area of clays as determined by ment (Hill et al., 1977). Thus, it is not only the
traditional methods of measurement. Thus, the specific reaction rate but also the area in contact
surface area of montmorillonite as determined by with the acid that determines the rate of removal of
nitrogen (N2) adsorption may be as high as 5 × 105 a specific mineral. An example calculation of rela-
m2/kg, whereas the reactive surface area is approxi- tive reaction rates of sandstone minerals is in
mately 104 m2/kg. The surface areas in Eq. 16-8 Sidebar 16B.
must be the reactive areas that are actually in con-

Table 16-5. Reaction rate constants for Eq. 16-11.

Mineral Ef K β α

 
 kg-mol minera l 
    kg-mol HF  −β 
α
 2  kg-mol HF      
 
 m - s    m3
 m3  

Potassium feldspar†
(orthoclase) 0.127 exp  − 4680  5.66 × 10–2 exp  956  0.4 1.2
 T   T 

Sodium feldspar†
(albite) 9.50 × 10–3 exp  − 3930  6.24 × 10–2 exp  554  1.0 1.0
 T   T 

α-quartz‡ 1.39 × 10–7 exp  − 1150  0 – –


 T 

Montmorillonite§ 1.1 × 10–2 exp  − 5200  0 – 1.0


 T 

Kaolinite§ 0.86 exp  − 6800  0 – 1.0


 T 


Fogler et al. (1973)

Adapted from Hill et al. (1977)
§
Adapted from Kline and Fogler (1981)

Reservoir Stimulation 16-7


16B. Relative reaction rates of sandstone minerals

A matrix acidizing treatment is aimed at overcoming the effects of near-wellbore formation damage. Ideally, the injected acid attacks
only the material causing the damage, which in most instances is clay particles or other fines. How efficiently the acid is being used
can be determined by calculating the reaction rates of all major mineral species present with the injected acid.
Consider a sandstone formation that has been damaged by the invasion of bentonite (montmorillonite) particles from drilling mud.
After the carbonate minerals have been removed by an HCl preflush, this clean sandstone contains 90% quartz, 5% albite (sodium
feldspar) and 5% montmorillonite by weight. The reactive surface areas of the minerals are 10 m2/kg for the quartz and albite and
8000 m2/kg for the montmorillonite. (A cube of quartz with a side of 1 mm has a surface area of 2.2 m2/kg; if it is 0.1 mm on a side, its
surface area is 22 m2/kg. Clays have a much larger surface area than detrital grains of quartz or feldspar.) Assume stoichiometric
ratios of 6 moles HF/mole quartz, 20 moles HF/mole feldspar and 40 moles HF/mole montmorillonite.
If this rock is contacted with a 12% HCl–3% HF solution at 125°F [50°C], what proportion of the HF will initially be consumed by
each of the three minerals? What are the mineral proportions of the rock dissolved?
Solution
Per unit mass of rock, the surface area of each mineral is its reactive surface area times the mass fraction of the mineral present in
the sandstone. For example, the reactive surface area of quartz per mass of sandstone Sq is (10 m2/kg)(0.9) = 9 m2/kg rock. Similarly,
the surface areas of feldspar and montmorillonite are 0.5 and 400 m2/kg rock, respectively. The acid concentrations in Eq. 16-11 are in
units of kg-mol/m3 solution (equivalent to gmol/L); the concentrations given as mass fractions are converted to these units by multiply-
ing by the solution density and the acid molecular weight, yielding 1.61 kg-mol HF/m3 solution and 3.53 kg-mol HCl/m3 solution. Quartz
is used to illustrate the calculation sequence to determine the reaction rates for each mineral.
First, the rate constant is calculated with the data from Table 16-5:

 1150  kg-mol quartz


E f = 1.39 × 10 −7 exp  −  = 3.95 × 10
−9
.
 273 + 50   kg-mol HF 
m2-s  3  (16B-1)
 m solution 
Then, the specific reaction rate for quartz from Eq. 16-11 is

kg-mol quartz
−rq = 6.59 × 10 −10 (1.61) = 1.06 × 10 −9 . (16B-2)
m 2-s

The overall reaction rate for quartz is the specific reaction rate multiplied by the reactive surface area:

kg-mol quartz
−R q = 1.06 × 10 −9 (9) = 9.54 × 10 −9 , (16B-3)
kg rock-s

which is multiplied by the molecular weight of quartz to put it on a mass basis:

kg quartz
−R q = 9.54 × 10 −9 (60.1) = 5.73 × 10 −7 . (16B-4)
kg rock-s

Finally, the rate of consumption of HF by the quartz reaction is obtained with Eq. 16-9, assuming 6 moles of HF are consumed for
each mole of quartz dissolved:

kg-mol HF
−R HF ,q = 9.54 × 10 −9 (6) = 3.44 × 10 −7 . (16B-5)
kg rock-s

The results of these calculations for all three minerals are summarized in Table 16B-1.
The fraction of HF expended in a particular reaction is the overall reaction rate for the mineral divided by the sum of the reaction
rates, which shows that 1.1% of the HF is reacting with quartz, 5.7% is reacting with feldspar and 93.2% is reacting with montmoril-
lonite. On the basis of the mass of mineral being dissolved, 95.1% of the rock dissolved is clay, 4.3% is feldspar, and less than 0.6% is
quartz. This is because of the high surface area of the authogenic clays (including, however, clay particles from drilling muds) and the
low reactivity of the quartz. Because clay and feldspar have relatively high reaction rates and generally form a small portion of the total
rock mass, they are dissolved first in sandstone acidizing. The quartz reaction becomes important in regions where most of the clay,
except clastic clays, and feldspar have already been removed.

Table 16B-1. Relative reaction rates of sandstone minerals.

Mineral –ri Si –Ri –RHF,i


(kg-mol i /m2-s) (m2/kg rock) (kg i /kg rock-s) (kg-mol HF/kg rock-s)

Quartz 5.73 × 10–8 9 3.44 × 10–6 3.44 × 10–7


Feldspar 1.77 × 10–7 0.5 2.32 × 10–5 1.77 × 10–6
Clay 1.81 × 10 –9
400 5.19 × 10 –4
2.89 × 10–5

Note: The subscript i denotes the mineral.

16-8 Fundamentals of Acid Stimulation


16B. Relative reaction rates of sandstone minerals (continued)

Notes for Sidebar 16B


Given 90% quartz, 5% albite and 5% montmorillonite by weight, the surface areas are 10 m2/kg quartz, 10 m2/kg albite and
8000 m2/kg montmorillonite. The reaction rates used are from Table 16-5.

• Quartz (MW = 60.1)


Because 1 kg rock has 0.9-kg quartz, the quartz area is (0.9)(10)90 m2/kg rock.

 1150 
E f = 1.39 × 10 −7 exp −  = 3.95 × 10
−9

 273 + 50 

kg-mol quartz
−rq = E f C HF = 6.36 × 10 −9
m 2-s

−R q = (6.36 × 10 −9 ) (9) = 5.72 × 10 −8


kg-mol quartz
kg rock-s

−R q = (5.72 × 10 −8 ) (60.1) = 3.44 × 10 −6


kg quartz
kg rock-s

 5.72 × 10 −8 kg-mol quartz  mol HF


R HF =   (6) = 3.43 × 10
−7

 kg rock-s  kg rock-s

• Albite (MW = 262, β = 1.0, α = 1.0)


 554 
K = 6.24 × 10 −2 exp  = 0.3468
 323 

CHCl = 3.53 kg mol/m3

 3930 
E f = 9.5 × 10 −3 exp −  = 4.938 × 10
−8

 323 

 103 m3   0.05 kg albite   −7 kg-mol albite  −8 kg-mol albite


RA =    1.77 × 10  = 8.84 × 10
 kg albite   kg rock   m2-s  kg rock-s

kg albite
R A = 8.84 × 10 −8 (262) = 2.32 × 10 −5
kg rock-s

kg-mol albite  kg-mol HF  −6 kg-mol albite


R A = 8.84 × 10 −8  20  = 1.77 × 10
kg rock-s  kg-mol albite  kg rock-s

• Montmorillonite (MW = 720)

 5200 
E f = 1.1× 10 −2 exp −  = 1.12 × 10
−9

 323 

−rmontmorillonite = (1.12 × 10 −9 )(1.61) = 1.81× 10 −9


kg-mol montmorillonite
m2-s

 m2   0.05 kg montmorillonite  m2
surface area =  8000   = 400
 kg montmorillonite   kg rock  kg rock

kg-mol montorillonite
R montmorillonite = 7.22 × 10 −7
kg rock-s

kg montorillonite
R montmorillonite = 5.20 × 10 −4
kg rock-s

kg montorillonite
R montmorillonite = 7.22 × 10 −7 (40) = 2.89 × 10 −5
kg rock-s

Reservoir Stimulation 16-9


• Reactions of fluosilicic acid with sandstone minerals rates, so that the potential precipitation zone is rapidly
As discussed in Section 16-2.1, fluosilicic acid is displaced away from the wellbore. Also, spent acid
produced when HF dissolves silicate minerals, and should be produced back immediately after the com-
the fluosilicic acid itself may then react with alumi- pletion of injection, because shutting in the well for
nosilicates. From models of coreflood experiments, even a relatively short time may allow significant sil-
Bryant (1991) and da Motta et al. (1992a, 1992b) ica precipitation to occur in the near-well vicinity.
suggested that the reaction between fluosilicic acid When ferric ions (Fe3+) are present, they can precip-
and clays and feldspars is slow at room tempera- itate from spent acid solutions as Fe(OH)3 when the
ture, but that it is of the same order of magnitude pH is greater than about 2. Ferric ions may result from
as the HF reactions with these minerals at tempera- the dissolution of iron-bearing minerals in an oxida-
tures above 125°F [50°C]. These conclusions have tive environment or may derive from the dissolution
been substantiated by more direct experimentation of rust in the tubing by the acid solution. When a high
(see Shuchart and Gdanski, 1996). level of ferric irons is likely in the spent acid solution,
sequestering agents can be added to the acid solution
to prevent the precipitation of Fe(OH)3. However,
16-2.3. Precipitation of reaction products Smith et al. (1969) suggested using these sequestrants
with caution, as they may cause more damage through
A major concern in acidizing, particularly the acidiz-
their own precipitation than would have been caused
ing of sandstones, is damage caused by the precipita-
by the iron.
tion of acid-mineral reaction products. In acidizing
Finally, in some reservoirs, contact of the crude
sandstones with HF, the formation of some precipi-
oil by acid can cause the formation of asphaltenic
tates is probably unavoidable. However, the amount
sludges. Simple bottle tests in which a sample of
of damage they cause to the well productivity depends
crude oil is mixed with the acid can indicate whether
on the amount and location of the precipitates. These
the crude has a tendency for sludge formation when
factors can be controlled to some extent with proper
contacted by acid. When sludge formation is a prob-
job design.
lem, emulsions of acid in aromatic solvents or sur-
The most common damaging precipitates that may
face-active additives have been used to prevent
occur in sandstone acidizing are calcium fluoride
asphaltene precipitation (Moore et al., 1965).
(CaF2), colloidal silica (Si(OH)4), ferric hydroxide
The tendency for precipitation reactions to occur in
(Fe(OH)3) and asphaltene sludges. Calcium fluoride
acidizing is predicted with comprehensive geochemical
is usually the result of the reaction of calcite with HF,
models of the chemical reactions between aqueous
according to
species and the host of minerals present. The most
CaCO3 + 2HF CaF2 + H2O + CO2 common type of geochemical model used to study
sandstone acidizing is the local equilibrium model,
Calcium fluoride is highly insoluble, so the precipi-
such as described by Walsh et al. (1982) and Faber et
tation of CaF2 is likely if any calcite is available to
al. (1994). This type of model assumes that all reac-
react with the HF. Inclusion of an adequate HCl pre-
tions are in local equilibrium; i.e., all reaction rates are
flush ahead of the HCl-HF stage prevents the forma-
infinitely fast. A typical result from this model is
tion of CaF2.
shown in Fig. 16-2, a time-distance diagram for the
Production of some colloidal silica precipitate is
injection of 11% HCl–4% HF into a formation con-
probably unavoidable in sandstone acidizing. The
taining calcite, kaolinite and quartz. This plot shows
equilibrium calculations of Walsh et al. (1982) show
regions where amorphous silica and aluminum fluoride
that there are virtually always regions where the spent
will tend to precipitate. A vertical line on the plot rep-
acid solution has the tendency to precipitate colloidal
resents the mineral species present as a function of dis-
silica. However, laboratory corefloods suggest that the
tance if all reactions are in local equilibrium. By
precipitation is not instantaneous and in fact may
coupling this model with a model of the formation per-
occur at a fairly slow rate (Shaughnessy and Kunze,
meability response to both dissolution and precipita-
1981) that, however, increases with temperature. To
tion, predictions of the productivity improvement
minimize the damage caused by colloidal silica, it is
expected from particular acid formulations may be
probably advantageous to inject at relatively high
obtained, as illustrated in Fig. 16-3 (Faber et al., 1994).

16-10 Fundamentals of Acid Stimulation


1.0 6

I
II
5

2.4
2.2
0.8 Productivity

1.4
III

2.6

2.6
Kaolinite improvement
and factor contours
calcite

HF concentration (%)
4
Fractional distance

1.8
0.6
Si(OH)4
AIF3 3
IV
Optimum
0.4 Si(OH)4
2
2.6
V 1.6
0.2

1.
1 2.2

2.
Dissolved zone

4
1.4

2.0
0 0
0 2 4 6 8 10 0 5 10 15 20 25
Pore volumes injected HCl concentration (%)

Figure 16-2. Time-distance diagram showing regions of Figure 16-3. Productivity improvement plot (Faber et al.,
possible precipitation (Schechter, 1992). 1994).

Recently, Sevougian et al. (1992) and Quinn (1994) damage than a local equilibrium model because the
presented a geochemical model that includes kinetics finite rate of the reactions allows displacing the pre-
for both dissolution and precipitation reactions (see cipitate farther from the wellbore.
Sidebar 16C). This model predicts less permeability

16C. Geochemical model predictions

An example presented by Quinn (1994) illustrates how acid formulation can be evaluated with a comprehensive geochemical model.
A high-quartz-content sandstone will be acidized with 100 gal/ft of 12% HCl–3% HF solution (commonly referred to as full-strength mud
acid). The mineralogy is illustrated in Fig. 16C-1. The region to be studied includes a damaged zone extending 6 in. beyond the wellbore.
Figure 16C-1.
High quartz (>80 vol%) Low clay (<5 vol%), low feldspar (<10 vol%) A representative
sandstone used in
acidizing simulation.

Microcline 6%
Quartz 85%
Albite 3%
Calcite 2%
Kaolinite 4%

The geochemical model predicts the distribution of acid and minerals after injection of the acid, as shown in Fig. 16C-2. All the HF
is consumed near the wellbore; some precipitation occurs, but the amorphous silica precipitation occurs beyond the damage zone,
where its effect is small. From these results, the porosity distribution around the wellbore is determined. Then, a model of the perme-
ability response generates a prediction of the productivity improvement expected for this treatment.

Reservoir Stimulation 16-11


16C. Geochemical model predictions (continued)
By repeating this modeling procedure for many different acid concentrations, the optimal acid formulation can be determined
(Fig. 16C-3). For this formation, 3% HF and high HCl concentrations are optimal.

Figure 16-C-2. Partial local equilibrium


Productivity improvement factor = 1.77
1.0 assumption (PLEA) model mineral profile
Reservoir minerals
of a high-quartz sandstone after acidiza-
Concentration (g-mol/L bulk volume)

0.9
0.8 tion with 100 gal/ft of 12% HCl–3% HF
0.7 injected at 0.1 bbl/min/ft at 125°F [50°C].
0.6 Kaolinite
0.5 Kaolinite (D)
Sodium feldspar
0.4
Potassium feldspar
0.3
0.2
0.1
0
0.25 0.75 1.25 1.75 2.25 2.75 3.25 3.75 4.25

3.0
Precipitates
Concentration (g-mol/L bulk volume)

2.5

2.0
Colloidal silica
1.5 AIF3 (s)
Na2SiF6
1.0 K2SiF6

0.5

0
0.25 0.75 1.25 1.75 2.25 2.75 3.25 3.75 4.25

4.5
Acid concentration
Concentration (g-mol/L bulk volume)

4.0
3.5
H+
3.0 HF
2.5
2.0
1.5 Original
damaged
1.0 zone
0.5
0
0.25 0.75 1.25 1.75 2.25 2.75 3.25 3.75 4.25
Radial distance (ft)

Figure 16C-3. Productivity improvement plot.


Volume of Reservoir mineralogy Injection
acid injected rate
High-quartz sandstone
100 gal/ft at 0.1 bbl/min/ft
6.0

1.8
1.5
HF concentration (%)

4.5 1.7 Productivity


1.6 improvement
factor (PIF)
PIF contours
1.8
1.7
1.5 1.6
1.5 1.5
1.4 1.4
1.3 1.3
0 1.2 1.2
0 6 12 18 24
HCl concentration (%)

16-12 Fundamentals of Acid Stimulation


16-3. Sandstone acidizing (Brannon et al., 1987). The benefits of lower concen-
tration HF solutions are a reduction in damaging pre-
16-3.1. Introduction cipitates from the spent acid and lessened risk of
unconsolidation of the formation around the wellbore.
A typical acid treatment in sandstones consists of the The selection of acidizing fluids should always begin
injection of an HCl preflush, with 50 gal/ft of forma- with an assessment of the formation damage
tion a common preflush volume, followed by the present—in general, the damaging material must be
injection of 50 to 200 gal/ft of HCl-HF mixture. A soluble in the treating fluids. Geochemical models can
postflush of diesel, brine or HCl then displaces the be used to guide acid selection, once the composition
HCl-HF from the tubing or wellbore. Once the treat- of the damaged formation is determined, as described
ment is completed, the spent acid should be immedi- in Section 16-2.3. Chapters 17 and 18 provide a com-
ately produced back to minimize damage by the prehensive treatment of acid selection for sandstone
precipitation of reaction products. and carbonate reservoirs, respectively.
A sandstone acidizing treatment design begins with
the selection of the type and concentration of acid to
be used. The volumes of preflush, HCl-HF mixture and 16-3.3. Sandstone acidizing models
postflush required and the desired injection rate(s) are
considered next. In virtually all acid treatments, the • Two-mineral model
placement of the acid is an important issue—a strategy Numerous efforts have been made over the years
to ensure that sufficient volumes of acid contact all to develop a comprehensive model of the sandstone
productive parts of the formation should be carefully acidizing process that could then be used as a
planned. Proper execution of the treatment is critical design aid. The most common model in use today is
to acidizing success, so the conduct of the treatment, the two-mineral model (Hill et al., 1977; Hekim et
including the mechanical arrangements for introducing al., 1982; Taha et al., 1989) that divides all minerals
the acid to the formation and the methods of treatment into two categories—fast-reacting and slow-reacting
monitoring, should be planned in detail. Finally, numer- species. Schechter (1992) categorizes feldspars,
ous additives are incorporated with acid solutions for authogenic clays and amorphous silica as fast react-
various purposes. The types and amounts of additives ing, and detrital clay particles and quartz grains are
to be used in the treatment must be determined on the the primary slow-reacting minerals. The model con-
basis of the completion, formation and reservoir fluids. sists of material balances applied to the HF acid and
These design factors are considered in detail in other reactive minerals, which for linear flow, such as in a
chapters: acid selection in Chapters 13 and 18, treat- coreflood, are
ment design (rate and volume) in Chapter 18, fluid δ(φCHF ) δC
placement and diversion in Chapter 19, treatment moni-
δt δx
{ }
+ u HF = − SF* VF E f ,F + SS*VS E f ,S (1 − φ)CHF
toring and evaluation in Chapter 20 and acid additives
(16-12)
in Chapter 15. This section presents models of the
sandstone acidizing process that provide a foundation
δ − MWHF SF* VFβ F E f ,F CHF
for the design methods used for field application.
δt
[ ]
(1 − φ)VF =
ρF
(16-13)

16-3.2. Acid selection δ − MWHF SS*VSβ S E f ,S CHF


The type and strength (i.e., concentration) of acid used δt
[ ]
(1 − φ)VS =
ρS
. (16-14)
in sandstones are selected primarily on the basis of
field experience with particular formations. For years, In these equations, CHF is the concentration of
the standard sandstone acidizing formulation consisted HF in solution and MWHF is its molecular weight,
of a 12% HCl–3% HF mixture, preceded by a 15% u is the acid flux, s is the distance, SF* and SS* are
HCl preflush. In fact, the 12% HCl–3% HF mixture the specific surface areas per unit volume of solids,
has been so common that it is referred to generically VF and VS are the volume fractions, Ef,F and Ef,S are
as mud acid. In recent years, however, the trend has the reaction rate constants (based on the rate of
been toward the use of lower strength HF solutions consumption of HF), MWF and MWS are the molec-

Reservoir Stimulation 16-13


ular weights, βF and βS are the dissolving powers a unit volume of rock pore space to the amount of
of 100% HF, and ρF and ρS are the densities of the mineral present in the unit volume of rock, which
fast- and slow-reacting minerals, respectively, for the fast-reacting mineral is
denoted by the subscripts F and S. When made
φ oβ F CHF
o
MWHF
dimensionless, assuming porosity remains constant, Ac( F ) = . (16-24)
these equations become ( o ) FoρF
1 − φ V

δψ δψ The Damköhler and acid capacity numbers for


+ + {Da ( F ) Λ F + Da ( S ) Λ S }ψ = 0 (16-15) the slow-reacting minerals are similarly defined.
δθ δε As acid is injected into a sandstone, a reaction
δΛ F front is established by the reaction between the HF
= − Da ( F ) Ac( F ) ψΛ F (16-16) and the fast-reacting minerals. The shape of this
δθ front depends on Da (F ). For low values of Da, the
convection rate is high relative to the reaction rate
δΛ S
= − Da ( S ) Ac( S ) ψΛ S , (16-17) and the front is diffuse. With a high Da, the reaction
δθ front is relatively sharp because the reaction rate
where the dimensionless variables are defined as is high compared with the convection rate. Figure
16-4 (da Motta et al., 1992a) shows typical concen-
CHF
ψ= o (16-18) tration profiles for high and low values of Da (F ).
CHF

VF 1.0
ΛF = (16-19) Ψ
VFo 0.9
(F)
ΛF
Da = 15
0.8 Da(S) = 0.43
V θ = 100
Λ S = So (16-20) 0.7 A c(F) = 0.006
VS A c(S) = 0
0.6
Ψ or ΛF

x 0.5 Da (F) = 0.4


ε= (16-21)
L 0.4

ut 0.3
θ= , (16-22)
φo L 0.2
0.1
where ψ is the dimensionless HF concentration, 0
Λ is the dimensionless mineral composition, ε is 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
dimensionless distance, θ is dimensionless time ε

(pore volumes), and φ is the porosity. The super- Figure 16-4. Acid and fast-reacting mineral concentration
script o denotes initial values prior to acid treat- profiles (da Motta et al., 1992a).
ment. For a coreflood, L is the core length. In
Eqs. 16-15 through 16-17, two dimensionless Equations 16-15 through 16-17 can be solved only
groups appear for each mineral: the Damköhler numerically in their general form. Numerical models
number Da and the acid capacity number Ac. These providing solutions to these equations, such as that
two groups describe the kinetics and the stoichio- presented by Taha et al. (1989), are frequently used
metry of the HF-mineral reactions. The Damköhler for acidizing design. However, analytical solutions
number is the ratio of the rate of acid consumption are possible for certain simplified situations.
to the rate of acid convection, which for the fast- Schechter (1992) presented an approximate solution
reacting mineral is that is valid for relatively a high Da (Da (F) > 10).

Da (F)
=
(1 − φ )V
o F
.
o
E (f F ) SF* L
(16-23)
This solution approximates the HF/fast-reacting-
mineral front as a sharp front, behind which all the
u
fast-reacting minerals have been removed. Conversely,
The acid capacity number is the ratio of the ahead of the front, no dissolution has occurred. The
amount of mineral dissolved by the acid occupying reaction between slow-reacting minerals and HF

16-14 Fundamentals of Acid Stimulation


behind the front serves to diminish the HF concen-
8
tration reaching the front. The location of the front is
7 Fast-reacting minerals present

θ=
(
exp Da ε f − 1(S)
) +εf, (16-25)
6 Reaction
(F) (S) front
A Dac 5
Fast-reacting minerals removed

x (cm)
which relates dimensionless time (or equivalently 4

acid volume) to the dimensionless position of the 3


front ε f, defined as the position of the front divided 2
by the core length for linear flow. The dimension- 1
less acid concentration behind the front is Perforation tunnel
0
ψ = exp( − Da ε ). (S) 0 4 8 12 16 20 24 28
(16-26)
z (cm)

A particularly convenient feature of this approxi- Figure 16-5. Ellipsoidal flow around a perforation
mation is that it is applicable to linear, radial and (Schechter, 1992).
ellipsoidal flow fields with the appropriate definition
of dimensionless variables and groups. Radial flow
These two positions should be sufficient for design
represents the flow of acid from an openhole, gravel-
purposes; the reader is referred to Schechter (1992)
pack or slotted liner completion and may also be a
for methods to calculate the complete acid penetra-
reasonable approximation of the flow from a perfo-
tion profile in this geometry (see Sidebar 16D).
rated well with sufficient perforation density. The
The characteristic lengths referred to in Table 16-6
ellipsoidal flow geometry approximates the flow
are the length of a core L, wellbore radius rw and
around a perforation (Fig. 16-5). The proper dimen-
length of the perforation lp. Different measures of
sionless variables and groups for these three flow
acid flow are used in which u is the linear flux in a
fields are given in Table 16-6. For the perforation
core, qi /h is the volumetric rate of acid injection per
geometry, the position of the front ε f depends on
foot into an openhole, and qperf is the volume of acid
position along the perforation. In Table 16-6, expres-
per time entering a perforation. The definition of Da
sions are given for the front position of the acid
must correspond to the geometry considered, but Ψ,
extending directly from the tip of the perforation
ΛF and Ac(F) as defined by Eqs. 16-18, 16-19 and
and for acid penetration along the wellbore wall.
16-24, respectively, apply to all geometries.

Table 16-6. Dimensionless groups in sandstone acidizing models.

Flow geometry ε θ Da (S )

Linear x ut (1− φ )V
o S
o
E f(S )SS* L
L φoL u

Radial r2
−1
q it (1− φ )V
o S
o
E f(S )SS* πrw2h
rw2 πrw2hφo qi

Ellipsoidal
Penetration from the tip of the perforation
1 3 2 z q perf t 2π(1− φo )l p3SS*VSoE f(S )
z −z + ; z =
3 3 lp 2πl φo
3
p
q perf

Penetration adjacent to the wellbore


3
1 1  1 x
x+  − 3; x = l
3  x + x 2 + 1 p

Reservoir Stimulation 16-15


16D. Comparison of acid volumes for radial 350
and perforation flow
300
The amount of acid required to remove damage beyond the Perforation
tip of a perforation is typically larger than the amount required 250

Acid volume (gal/ft)


to remove damage to the same distance in radial flow. This is
illustrated by using Schechter’s (1992) model to find the vol- 200
ume of acid required to penetrate a given distance for these
two geometries. 150
Consider a formation with 20% porosity containing 10%
fast-reacting minerals (feldspar, clay or both), 5% calcium car- 100
bonate and 85% quartz. A damage region extends 6 in. into
the formation (a radial region or 6 in. beyond the tip of a per- Radial
50
foration), and the initial skin effect resulting from the damage
is 10. The wellbore radius is 0.328 ft. In the perforated well 0
case, there are 4 shots per foot (spf) and the perforations are 0 1 2 3 4 5 6
6 in. long with a diameter at the wellbore of 0.5 in. A 12%
HCl–3% HF solution is injected at 0.1 bbl/min/ft after the injec- Acid penetration distance (in.)
tion of sufficient preflush of 15% HCl to remove the carbon-
ates from the region to be contacted with live HF. Determine Figure 16D-1. Acid penetration for radial and perforation
the acid volume required as a function of the penetration dis- flow.
tance of the acid and the skin effect evolution for radial and
perforation geometries. The downhole treating temperature
is 125°F [50°C]. For these conditions, the Damköhler number
Da for the slow-reacting mineral in radial flow is 0.013. 5

Solution 4
Equation 16-25 can be used with the appropriate definitions
of the dimensionless variables and groups for the two geome-
3
tries from Table 16-6. The acid capacity number Ac is the
Perforation
same for either geometry; the ratio of the Damköhler numbers
Skin effect

is 2
3

(2) 12 
6
4
Da perf 213perf (SPF ) 1
= = = 93, (16D-1)
(0.328)
2 2 Radial
Darad rw

0
where SPF is the perforation density in spf. Da for the slow-
mineral reaction is calculated as 0.12 for perforation flow. Ac
for the fast-mineral reaction is 0.021. (The values of Da and –1
Ac used in this example were obtained from laboratory core- 0 100 200 300 400 500
flood tests as described by Economides et al., 1994).
Using these values in Eq. 16-25 for acid penetration rang- Acid volume (gal/ft)
ing from 0 to 6 in. obtains the results shown in Figs. 16D-1
and 16D-2. For acid penetrations beyond 2 in., more acid is Figure 16D-2. Reduction in skin effect value for radial
required for the perforation geometry than for radial flow. The and perforation flow.
skin effect evolution reflects the larger volumes of acid
required to penetrate through the damaged region for the
perforation geometry compared with the radial geometry.

It is interesting to note that the slow-reacting min- is assumed, implying that Da(F) is infinite. This solu-
eral Da and the fast-reacting mineral Ac are the only tion can be used to estimate the volume of acid
dimensionless groups that appear in this solution. required to remove the fast-reacting minerals from
Da(S) regulates how much live HF reaches the front; a given region around a wellbore or perforation.
if the slow mineral reacts fast relative to the convec- The dimensionless groups Da(S) and Ac(F) can be
tion rate, little acid is available to propagate the fast- calculated with Eqs. 16-23 and 16-24, respectively,
mineral front. Ac for the slow-reacting mineral is not and Table 16-6 on the basis of the rock mineralogy
important because the supply of slow-reacting min- or can be obtained from experiments.
eral is almost constant behind the front. Ac(F) directly
• Two-acid, three-mineral model
affects the frontal propagation rate—the more fast-
reacting mineral present, the slower the front will Recently, Bryant (1991) and da Motta et al. (1992b)
move. Da(F) does not appear because a sharp front presented evidence that the sandstone acidizing

16-16 Fundamentals of Acid Stimulation


process is not well described by the two-mineral the fast-reacting minerals with a given volume of
model, particularly at elevated temperatures. These acid because the fluosilicic acid also reacts with
studies suggest that the reaction of fluosilicic acid these minerals and the reaction product of silica gel
with aluminosilicate (fast-reacting) minerals may (Si(OH)4) precipitates. This reaction allows live
be significant. Thus, an additional acid and mineral HF to penetrate farther into the formation; however,
must be considered to accommodate the following there is an added risk of a possibly damaging precip-
reaction, which is added to the two-mineral model: itate forming.
Sumotarto (1995) presented an example that illus-
H2SiF6 + fast-reacting mineral →
trates the improved performance predicted with the
νSi(OH)4 + Al fluorides
two-acid, three-mineral model compared with the
The practical implications of the significance of one-acid, two-mineral model. Figure 16-6 compares
this reaction are that less HF is required to consume the mineral concentration profiles predicted by these

One-acid, two-mineral model Two-acid, three-mineral model

1.2 Injection volume = 25 gal/ft 1.2 Injection volume = 25 gal/ft

1.0 1.0
0.8 0.8
0.6 0.6
C/Co

C/Co
0.4 0.4
0.2 0.2
0 0 Mineral 1
Mineral 1 Mineral 2
–0.2 Mineral 2 –0.2 Mineral 3
0 2 4 6 8 10 12 0 2 4 6 8 10 12
Distance (in.) Distance (in.)

1.2 Injection volume = 100 gal/ft 1.2 Injection volume = 100 gal/ft

1.0 1.0
0.8 0.8
0.6 0.6
C/Co

C/Co

0.4 0.4
0.2 0.2
0 0 Mineral 1
Mineral 1 Mineral 2
–0.2 Mineral 2 –0.2 Mineral 3
0 2 4 6 8 10 12 0 2 4 6 8 10 12
Distance (in.) Distance (in.)

1.2 Injection volume = 225 gal/ft 1.2 Injection volume = 225 gal/ft

1.0 1.0
0.8 0.8
0.6 0.6
C/Co

C/Co

0.4 0.4
0.2 0.2
0 0 Mineral 1
Mineral 1 Mineral 2
–0.2 Mineral 2 –0.2 Mineral 3
0 2 4 6 8 10 12 0 2 4 6 8 10 12
Distance (in.) Distance (in.)

Figure 16-6. Dimensionless mineral concentrations at various injection volumes.

Reservoir Stimulation 16-17


two models for the injection of 12% HCl–3% HF precipitation is predicted by the two-acid, three-min-
into a damaged formation composed initially of 17% eral model. Using a model of the permeability
clays and feldspars (fast reacting) and 83% quartz response to both mineral dissolution and precipita-
(slow reacting). In this figure, mineral 1 is clay and tion, the permeability and skin effect response are
feldspar, mineral 2 is quartz, and mineral 3 is silica predicted for each model and compared in Figs. 16-7
gel. After 100 gal/ft of injection, the feldspars and and 16-8. Although some precipitation is indicated
clays have been dissolved in a region extending by the two-acid, three-mineral model, improved per-
about 6 in. beyond the wellbore according to the formance because of the fluosilicic acid reactions is
two-acid, three-mineral model, whereas only 2 in. predicted compared with the one-acid, two-mineral
of dissolution is predicted by the one-acid, two-min- model.
eral model. In addition, a significant zone of silica

One-acid, two-mineral model 1000


1000 Model 1
25 gal/ft Model 2
100 gal/ft
225 gal/ft 800
800

Permeability (md)
Permeability (md)

600
600

400
400

200
200

0 0
0 2 4 6 8 10 12 0 100 200 300 400 500
Distance (in.) Injection volume (gal/ft)

6
Two-acid, three-mineral model Model 1
Model 2
1000 5
25 gal/ft
100 gal/ft
800 225 gal/ft 4
Skin effect
Permeability (md)

3
600

2
400
1
200
0
0 100 200 300 400 500
0
Injection volume (gal/ft)
0 2 4 6 8 10 12
Distance (in.)
Figure 16-8. Top: Average permeability versus injection
Figure 16-7. Permeability at various injection volumes. volume obtained from running the simulator with models 1
(one-acid, two-mineral model) and 2 (two-acid, three-mineral
model). Bottom: Skin effect calculated using models 1 and 2
as a function of injection volume (time).

16-18 Fundamentals of Acid Stimulation


16-3.4. Permeability response The Lambert expression is identical to that of Lund
and Fogler when M/∆φmax = 45.7.
To predict the response of a formation to acidizing, it is Using the values of the constants suggested, the
necessary to predict the change in permeability as acid Labrid correlation predicts the smallest permeability
dissolves some of the formation minerals and other increase, followed by the Lambert and then the Lund
minerals precipitate. The permeability change as a and Fogler correlations. The best approach in using
result of acidizing is an extremely complicated process these correlations is to select the empirical constants
because it is affected by several different, sometimes on the basis of coreflood responses, if available. If
competing phenomena in the porous media. The per- data are lacking for a particular formation, the Labrid
meability increases as the pores and pore throats are equation will yield the most conservative design.
enlarged by mineral dissolution. At the same time,
small particles are released as cementing material is
dissolved, and some of these particles lodge (perhaps
temporarily) in pore throats, reducing the permeability.
16-4. Carbonate acidizing
Any precipitates formed also tend to decrease the per-
16-4.1. Distinctive features
meability. The formation of carbon dioxide (CO2) as
carbonate minerals are dissolved may also cause a tem- In this chapter, sandstone acidizing is distinguished
porary reduction in the relative permeability to liquids. from carbonate acidizing although sedimentary rocks
The result of these competing effects is that the perme- exhibit a spectrum of compositions ranging from
ability in corefloods usually decreases initially; with almost pure calcite or dolomite to very clean sands.
continued acid injection, the permeability eventually The fundamental distinguishing feature is the HCl-
increases to values significantly higher than the original soluble fraction. If the HCl solubility of a rock is less
permeability. than 20%, a sandstone treatment using an HCl-HF
The complex nature of the permeability response mixture (for a discussion of such rules of thumb, see
has made its theoretical prediction for real sandstones McLeod, 1984) would most likely be applied. For-
impractical, though some success has been achieved mations composed largely of calcite or dolomite,
for more ideal systems such as sintered disks (Guin et including chalks and marls, are largely soluble in HCl
al., 1971). As a result, empirical correlations relating and are candidates for carbonate acidizing using HCl
the permeability increase to the porosity change dur- without HF.
ing acidizing are used. The most common correlations Carbonate acidizing with HCl is not complicated by
are those of Labrid (1975), Lund and Fogler (1976) a tendency for precipitates to form, as is the case for
and Lambert (1981). The Labrid correlation is sandstone acidization. As shown by the typical reac-
n
tions in Table 16-1, the reaction products CO2 and
k  φ CaCl2 are both quite water soluble (for a discussion of
= M  , (16-27)
ko  φo  their solubilities, see Shaughnessy and Kunze, 1981;
Schechter, 1992). Therefore, the formation of a precipi-
where ko and φo are the initial permeability and porosity
tate or a separate CO2-rich phase is generally not a
and k and φ are the permeability and porosity after
problem. Even if CaCl2 precipitates or a CO2 phase sep-
acidizing, respectively. M and n are empirical constants,
arates, these phases are readily dissolved when oil (or
reported to be 1 and 3, respectively, for Fontainebleau
gas) and water production is resumed. Despite the sim-
sandstone.
plified chemistry, HCl acidizing is a difficult process to
The Lund and Fogler correlation is
model. The origin of the difficulty is the rate at which
k   φ − φo   the reactions take place as compared with those of HF
= exp  M   , (16-28) with the various minerals prevalent in sandstones.
ko   ∆φ max  
Reaction rates are discussed in Section 16-2.2, and it
where M = 7.5 and the difference in maximum por- is instructive to compare some of them. HCl reactions
osity is ∆φmax = 0.08 from best-fit data for Phacoides with carbonates are orders of magnitude faster than HF
sandstone. reactions with sand (quartz), clays, etc.
The Lambert correlation is Because of the high reaction rate, HCl tends to etch
preferred pathways in carbonate rocks, apparently fol-
k
ko
[ ]
= exp 45.7(φ − φ o ) . (16-29) lowing local high-permeability streaks (Wang, 1993),

Reservoir Stimulation 16-19


rather than progressing through the formation as a uni-
form front, as is the case in sandstone acidizing. These
pathways are soon enlarged by acid reaction at the
walls into sizable holes that have a diameter much
larger than that of naturally occurring pores. The
process continues until a few large holes become so
dominant that essentially all the injected acid flows
through these pathways, both enlarging and extending
them.
It is this tendency for macroscopic pathways to form
that makes HCl acidizing difficult to model. Because
the holes that form are large, they become the most
important feature of the process. To model carbonate
acidizing, the formation of these holes must be taken
into account. In fact, it is believed that the success of
acid stimulation of carbonate formations is due to the
formation of these preferred flow paths extending out-
ward from the wellbore or perforation. If the pathways
extend through the damaged zone, the produced fluids
can flow into the wellbore through these flow paths
with relatively little pressure drop because the holes
are much larger than the natural pores.
Thus, the fundamental physics of carbonate acidiz-
ing is embodied in three topics discussed in this sec-
tion. The first concern is the characterization of the
holes or flow paths created by the acid. Second, the
conditions under which they form must be defined.
And third, the rate at which they are extended is an Figure 16-9. Wormholes created by acid dissolution of lime-
issue of considerable practical significance. stone (Hoefner and Fogler, 1988; reproduced with permis-
sion of the American Institute of Chemical Engineers.
Copyright 1998 AIChE. All rights reserved.).

16-4.2. Wormholes
It is not known who first described the acid-etched Daccord and Lenormand (1987) considered the
pathway as a “wormhole,” but this appellation is com- characterization of a wormhole in terms of its fractal
monly accepted by those familiar with the complex dimension. A fractal is a self-similar geometric pattern.
etch pattern produced by acidizing carbonate cores in This implies that under increasing magnification the
the laboratory. Perhaps it was A. R. Hendrickson of same pattern will continue to reappear. Thus, accord-
Dowell. Figure 16-9 shows that the terminology is apt. ing to this notion the structure of a large wormhole is
This is a photograph of a metal casting of a wormhole repeated with branches from the main trunk that are
created by forcing molten metal into a wormhole, smaller replicas of the larger one. This replication is
allowing the metal to solidify and then dissolving the repeated as the magnification is increased until the
remaining rock with HCl. The casting illustrates the pores of the native rock come into view. These do not
complex morphology of the etch pattern. This is typi- resemble acid etch patterns because they were created
cal of many castings, which have been produced under by different processes. The discovery by Daccord and
a variety of experimental conditions. The chaotic Lenormand that wormholes are fractals is a significant
nature of the pattern seemingly discourages any contribution.
attempt to characterize its structure. However, it has One manifestation of the repeating, or self-similar,
been suggested that there is an underlying regularity character is that the perimeter or the length of a worm-
that may be useful for modeling. hole increases as the degree of magnification used in

16-20 Fundamentals of Acid Stimulation


its measurement increases. This is because at higher flux entering the matrix is quite small, wormhole initi-
magnification, more of the detailed structure becomes ation is not prolific, thereby indicating a flow-rate
evident and is, therefore, susceptible to measurement. dependence of the initiation process. The proof of this
For self-similar systems, the length of a wormhole assertion is based primarily on metal casts of worm-
plotted against the length of a ruler used to measure holes, such as that shown by Fig. 16-9. However,
the length is a straight line on a log-log plot. The slope adopting the notion of a flow-rate-dependent initiation
of this line is related to the fractal dimension. In the process allows interpreting the results of laboratory
case of wormholes, Daccord and Lenormand reported acidizing experiments and understanding the origin of
that the fractal dimension is 1.6. This implies that the the fractals. Furthermore, this approach leads to pre-
length of a wormhole is proportional to L1.6 rather than diction of the optimum injection rate in linear core
L, where L is the macroscopic length. Daccord et al. experiments that has been experimentally observed.
(1989) utilized this fractal dimension in developing a Thus, the analysis presented here represents a founda-
field design method for carbonate acidizing. In their tion upon which the design of acid treatments can be
approach, the complexity of the wormhole is portrayed based, but further work is required to achieve the
by its fractal dimension. This is only a partial charac- desired goal, namely, the ability to predict the stimula-
terization, because it is not possible to describe the tion resulting from an acid treatment given the essen-
minute details of wormhole geometry. tial parameters of acid composition, injection rate,
Although a precise description of a wormhole is not formation temperature and rock properties.
attainable, it is desirable to have a model that provides The initiation of wormholes occurs when live acid
guidance in determining the best treatment parameters. penetrates into pores present in the native rock. These
What should be the injection rate? Should the injection pores are distributed in size and shape; therefore, the
rate be constant during the entire course of the treat- amount of acid flowing through each of the pores dif-
ment? What acid type and concentration are best? fers. The rate at which a given pore is enlarged by the
What additives should be included in the acidic solu- acid depends, of course, on the amount of acid enter-
tion? These questions relate to controllable variables ing that pore and the fraction of the acid reacted at the
and, therefore, must be addressed each time an acid walls of the pore before the acid exits and then enters
treatment is designed. Rather than depending solely on other pores located downstream. Thus, even at the
past experience, some theoretical help is welcome for pore level, the processes that contribute to the creation
developing the best strategy possible based on the of an etch pattern are complex, involving convection,
information available. diffusion and chemical reactions within each of the
The problem is approached in the following sec- invaded pores. It has not been proved practical to con-
tions by addressing two separate issues, both of which sider these processes in a single pore and then attempt
are relevant to the questions posed. The first concerns to consider the collective behavior to derive a macro-
the conditions requisite for the initiation of worm- scopic etch pattern. Schechter and Gidley (1969) used
holes, and the second deals with their growth or prop- this approach, but to make progress using their results
agation. Both of these studies provide information requires knowing in advance the entire distribution of
required for the design of carbonate acid treatments. pore sizes, permeability and porosity of the native
The Appendix to this chapter discusses advances rock to be acidized. Even armed with this knowledge,
in understanding and predicting wormhole formation. which is seldom available, prediction of the etch pat-
tern is not routine.
The prediction of wormhole initiation is, however,
16-4.3. Initiation of wormholes based on a result that emerged from considering the
The fractal, or self-similar, topology of a wormhole behavior of each pore in the medium. If a pore is rep-
structure implies that the mechanism for the initiation resented as a cylindrical hole with a radius R and a
of wormholes is a “local” phenomenon that occurs length l, then the rate at which the pore cross-sectional
continuously along its bounding surfaces as well as at area A increases as a result of acid reaction at the pore
its tip. Thus, tiny wormholes may be initiated when- walls may, in general, be written in the form
ever live acid enters the pores of the virgin rock irre- dA
spective of the etch pattern already in existence. = ψA n , (16-30)
dt
Experiments have shown that in cases where the acid

Reservoir Stimulation 16-21


where ψ is a function of the fluid velocity in the pore, Equation 16-32 looks formidable, but the goal is to
reaction rate and other parameters that determine the reduce this expression to a form that reveals the expo-
rate at which rock is dissolved. The exponent n is also nent n. To accomplish this goal, two extreme condi-
a function of the many parameters that contribute to tions are considered. First, examine the limit
ψ. The advantage of representing the rate of acid reac-
tion within a pore in this form is that the stability of 2 πE f C0m−1l
the enlargement process depends on the value of the << 1, (16-34)
Av
exponent n (see Schechter, 1992). If n > 1 for a certain
few pores in the native rock, then these pores become which implies that only a small fraction of the acid
larger faster than all the other pores that have growth reacts within the pore. In this limit, which applies for
characterized by exponents smaller than unity. This large v–, the rate of pore enlargement reduces to
criterion for uncontrolled growth determines whether dA
a wormhole is initiated. The growth of each individual → 2 π E f C0m−1 XA1/ 2 = ψ 1 A1/ 2 . (16-35)
dt
pore is characterized by the value of the exponent n.
Wormholes form whenever one or more of the pores Thus, in this limit the exponent n is 1⁄2 (n < 1) and
grow at a rate determined by n > 1. pores with a cross-sectional area such that the inequal-
Thus, this criterion for wormhole initiation hinges ity of Eq. 16-34 is satisfied do not form wormholes.
on the value on the exponent n. To investigate the fac- The pores will enlarge rather uniformly, and the acid
tors determining n, recall that the rate of acid reaction front will remain sharp, progressing through the porous
at a pore wall is given by the empirical expression matrix also rather uniformly. Not all pores, however,
satisfy the inequality. A second limit for some of the
reaction rate = − E f CWm . (16-31) pores may be possible. This second limit occurs when
the pores are of such a size that the inequality of
The acid concentration CW in this equation is the Eq. 16-34 is reversed, implying almost complete acid
concentration near the solid/liquid interface. The acid reaction within these pores. In this second limit, which
that reacts at the pore surface must be replenished by applies for small v–, can be found
acid diffusing from the bulk solution to the pore wall.
If this diffusion rate is slow, the rate of pore area dA vXC0 2
→ A = ψ 2 A2 . (16-36)
enlargement may be limited by diffusion. However, dt 8πl
the native pores are generally small enough so that
Thus, pore areas satisfying the second limit are
diffusion is relatively fast, and the average velocity
unstable (n > 1). They grow more rapidly than the
within a pore v– is sufficient to maintain a rate of mass
neighboring pores. They become small wormholes and
transfer to the surface so that the acid reaction at the
continue to evolve into the macroscopic etch patterns
wall is the controlling factor. The rate at which a pore
shown by the metal casts.
is enlarged is, therefore, (see Schechter, 1992)
For a given reaction rate and acid flux, there are
dA   2 π E f C0m−1l   pores that are essentially too small to become worm-
l = vAXC0 1 − exp −  , (16-32) holes and perhaps others that are of sufficient size to
dt   Av   exhibit uncontrolled growth and eventually become
where l is the length of a pore defined as the distance macroscopic. Thus, for a given acid flux, the native
the acid in a pore travels before mixing with acid pores may fall into two different categories: candidates
emerging from other pores. The acid concentration for incipient wormhole formation and noncandidates.
C0 is the concentration at the pore entrance. There is, therefore, a critical (or transitional) pore size
The average velocity in a pore v– depends on the AT that may be estimated as the value where the two
local Darcy flux u and the pore cross-sectional area. limiting growth rates become equal:
The flow within a single pore is laminar, so that ψ 1 AT1/ 2 = ψ 2 AT2 . (16-37)
dp A uA
v= = , (16-33) Solving this expression for AT yields
dt 8πµ 8πk
AT ≅ 20 Da 2 / 3 ( kl ) ,
2/3
(16-38)
where k is the permeability of the rock matrix.

16-22 Fundamentals of Acid Stimulation


where Da = EfC0m – 1/u and k is the formation perme- 16-4.4. Acidizing experiments
ability (Wang, 1993; Wang et al., 1993). Thus, if all
Essentially two different types of results are found
the pores in the native rock have cross-sectional areas
on the basis of acidizing carbonate cores in the labora-
less than AT, wormhole initiation cannot occur until at
tory. One is the metal cast of a wormhole and the
least one of the pores has been enlarged by acid reac-
other is the pressure drop measured while acidizing
tion to a size sufficient to allow wormhole develop-
at a constant injection rate. Both types of experiments
ment. The critical, or transitional, area depends on
have proved instructive. The transition area defined by
both the reaction rate and the acid flux. This condition
Eq. 16-39 depends on the acid flux and reaction rate,
has, as discussed subsequently in this chapter, consid-
which in turn is a function of the acid concentration,
erable practical relevance.
reaction temperature and rock composition. The acid
The average length of a pore is a rather nebulous
flux is the easiest to control and is the most widely
quantity that, on the basis of a number of laboratory
studied variable.
experiments using two different limestones and a
dolomite, appears to be about 0.1 mm. If we use this • Acid flux
value, the criterion for the critical pore dimension The flow rate is expected to influence the acid etch
becomes pattern for reasons that may be best understood by
AT = 0.93[ Da k ] , considering the idealized depiction of a wormhole
2/3
(16-39)
shown by Fig. 16-10. It is a cylinder with fluid loss
where both k and AT must be expressed as cm2.† about the perimeter as well as at the tip. Depending
This equation is quite simple in appearance, but its on the external pressure field surrounding the worm-
implications are profound and these may be tested hole, the fluid-loss flux may vary from point to point
experimentally. Laboratory results that are seemingly about the surface of the cylinder. If the flux into the
counterintuitive may be satisfactorily explained by rock is small at some points, AT as determined by
invoking the concept embodied by Eq. 16-39. Eq. 16-38 may exceed the cross-sectional area of all

Wormhole

Injection pressure, pinj Wormhole radius, rwh

Total injection rate, qT


Flow at the tip, qe

Fluid loss, qL
Injection rate at core face, qc

Length to tip, Le

Figure 16-10. Single-wormhole model.


Given a permeability in md, multiply by 9.869 × 10–12 to obtain
the dimensions in cm2.

Reservoir Stimulation 16-23


the native pores, and wormholes will not form at that rate. The left-hand core at the lowest flow rate is an
point. The walls of the wormhole will then be eroded example of a nearly uniform dissolution front, where
in a generally uniform manner. If, however, the flux the inlet face of the core has essentially been dis-
at the tip is large enough to initiate wormholes, a solved. The flow rate was evidently less everywhere
network of small wormholes will continuously form than the critical one. As predicted, the wormholes
at the tip, rapidly extending its length. that developed at the higher flow rates show substan-
In summary, it is expected that for injection rates tial branching, displaying the fractal structure postu-
that are very slow, wormholes will not form and the lated by Daccord and Lenormand (1987). At higher
face of the core will dissolve rather uniformly. At rates, much of the acid is expended in the creation
modest injection rates, large enough to initiate of the highly ramified structure shown by the casts.
wormholes at the tip of the primary wormhole, an Of primary practical interest are the wormholes
etch pattern is expected to develop that shows little created at intermediate injection rates. They develop
branching from the primary wormhole. Most of the a minimum of side branches extending from the
acid is then expended in extending the primary perimeter of the main channel, in agreement with
wormhole. If the rate of acid injection is then the etch pattern anticipated by consideration of a
increased, the acid fluid-loss flux into the rock critical pore area. Indeed, the series of casts in
matrix may be large enough everywhere—or at Fig. 16-11 confirms the existence of a transitional
least at many points—to allow the initiation of area. Although the casts provide strong evidence
wormholes along the boundary of the primary one. supporting the hypothesis set forth in the preceding
A highly ramified wormhole structure is expected section, they also suggest a means for further verifi-
at the higher injection rates. cation. It seems evident in considering the series
Hoefner and Fogler (1988) prepared metal casts of of casts that a wormhole formed with a minimum
wormholes that developed in calcite cores at various of side branching will penetrate through the core
rates of injection (Fig. 16-11). The casts are arrayed using a smaller quantity of acid than would be
from left to right to correspond to increasing flow required otherwise.

Figure 16-11. Metal casts of wormholes that developed in calcite cores at various rates of injection are arrayed from left to
right to correspond to increasing flow rate (Hoefner and Fogler, 1988).

16-24 Fundamentals of Acid Stimulation


Thus, Eq. 16-39 appears to be the key to the deter-
6
Indiana limestone mination of the effect of acid injection rates in field
Glen Rose limestone
operations. Ideally, a few dominant wormholes
5
1-in. diameter × 6-in. long should be initiated extending from the wellbore in
Pore volumes to breakthrough

15% HCl (4.4N)


Room temperature openhole completions or from the perforations in
4
cased holes. This would initially entail a modest acid
injection rate (calculated by Eq. 16-39). As the acid
3
treatment progresses and the lengths of the few dom-
inant wormholes increase, higher injection rates are
2
then desirable to continue to extend the dominant
wormholes, if possible. Nierode and Williams
1 (1971) were apparently the first to emphasize the pri-
mary importance of fluid loss from wormholes as a
0 limiting factor in extending them. Their hypothesis
0 100 200 300 400 500
Injection rate (mL/hr)
is in no way weakened by the experiments and
analysis presented here. Acid fluid loss from a
Figure 16-12. Coreflood results for Indiana and Glen Rose wormhole remains a limiting factor in the propaga-
limestones. tion of wormholes.
• Reaction temperature
Wang et al. (1993) measured the volume of acid Increasing the reaction temperature increases Ef
required to achieve wormhole penetration through exponentially (see Eq. 16-9). In accordance with
a core (breakthrough). Figure 16-12 is a plot of the Eq. 16-39, this should result in a corresponding
acid pore volumes to breakthrough as a function of increase in AT. Therefore, the optimum injection rate
the injection rate. As anticipated, an optimum must correspondingly be increased (see Sidebar 6E).
injection rate exists. Wang et al. calculated the acid This predicted trend is shown by Fig. 16-13. The
fluxes about the wormhole that develop during the optimum injection rate at a temperature of 125°F
experiment to find the flux at the tip as well as is almost twice as large as the optimum at room
along the sides of the wormhole. They calculated temperature. According to Wang et al. (1993), the
for the optimum case (i.e., the experiment requiring increased rate is predicted by Eq. 16-39. An impli-
the least volume of acid) that the flux at the tip is cation of this result is that, if possible, deep wells
well predicted by Eq. 16-39. The critical size AT should be acidized at higher rates than shallow
was determined by the capillary entry pressure and ones; but in either case, increasing the rate during
the reaction rate was determined by Eq. 16-8. the course of the treatment is apt to be beneficial in
A subsequent study by Bazin et al. (1995) using extending the wormholes. Initially, however, the
two other limestones (Lavoux and Savonnieres) acid flux should be restricted by the value deter-
shows the importance of core length in laboratory mined with Eq. 16-39 if this rate is possible without
studies. As the length of the dominant wormhole fracturing the formation. As shown by Figs. 16-12
increases, the amount of acid lost through the lateral and 16-13, the volume of acid required to achieve
boundaries increases, thereby reducing the volume breakthrough does not increase rapidly for injection
of acid reaching the tip and ultimately resulting in an rates in excess of the optimum, so maintaining the
acid flux at the tip that is too small to initiate worm- optimum injection rate is not thought to be critical.
holes there. In such a case, the wormhole can extend Stimulations conducted at rates somewhat in excess
only slowly while it is also being enlarged. Worm- of those demanded by Eq. 16-39 may not differ
hole growth is slowed, as observed by Bazin et al. greatly from those achieved at optimum. On the
T. Huang (pers. comm., 1996) studied the rates of other hand, the amount of acid to achieve break-
wormhole propagation reported by Bazin et al. and through does increase substantially for rates less
found them predictable on the basis of Eq. 16-39 than the optimum, so maintaining a sufficient rate
when fluid loss is properly taken into account. if possible is recommended.

Reservoir Stimulation 16-25


16E. Optimum injection rate for initiating 7
carbonate treatment Room temperature
125°F
6
Consider an acid treatment of a well in a carbonate formation Indiana limestone

Pore volumes to breakthrough


known to be composed of calcite with little dolomite and to 1-in. diameter × 6-in. long
5 3.4% HCl (1N)
have a permeability of 5 md. The treatment will be conducted
using 15% HCl injected initially, if possible, at a rate nearly
corresponding to the optimum one. The reservoir tempera- 4
ture is 125°F [50°C]. The thickness of the formation to be
treated is 30 ft. The well is completed openhole with a well-
bore diameter of 6 in. 3
To determine the optimum injection rate, the cross-
sectional area of the largest native pores must be estimated. 2
Generally, this cross-sectional area is determined from the
capillary entry pressure measured by mercury porosimetry, if
an adequate core sample is available. For the present exam- 1
ple, use AT = 1.45 × 10–7 cm2. If a sample of the formation is
not available, then AT may be estimated as a multiple of the 0
formation permeability (Dullien, 1979). What is the optimum 0 100 200 300 400 500
acid injection rate?
Injection rate (mL/hr)
Solution
Based on the data provided in Table 16-4, Figure 16-13. Indiana limestone coreflood results at differ-
ent temperatures.
 7.55 × 10 3 
E f = 7.291× 10 7 exp −  = 5.14 × 10 ,
−3
(16E-1)
 (50 + 273) 
ramified structure that is inefficient in developing an
and because the acid is 15%, C0 = 4.4 kg-mol/m3
(= 4.4 mol/L). Therefore,
etch pattern consisting of a few dominant wormholes.
It is, therefore, expected that corresponding to a sub-
E f C 0m −1 =
(5.14 × 10 )(4.4)
−3 0.63

= 2.97 × 10 −3 m/s. (16E-2) stantial decrease in the reaction rate, there must be
4.4 an associated decrease in the acid injection rate to
In an appropriate set of units, k = (5 md)9.869 × 10–12 remain at optimum. Figure 16-14 shows the acid
cm2/md = 4.9345 × 10–11 cm2. Therefore, the optimum flux pore volumes to breakthrough as a function of the
from Eq. 16-39 is u = 2.38 × 10–3 m/s = 0.0078 ft/s. This
corresponds to an initial acid injection rate of 3.9 bbl/min injection rate for dolomite cores. The optimum rate
(0.13 bbl/min/ft). If this injection rate exceeds the formation at room temperature is not readily discernible, but
parting pressure, injection would be at the highest possible
matrix rate during the entire treatment. it is evident that slow rates are preferred to higher
The optimum rate decreases with formation temperature. ones. Hoefner and Fogler (1988) also studied the
Consider the same treatment when the temperature is 85°F
[30°C] rather than 125°F. Then, Ef = 1.1 × 10–3. This leads
acidization of dolomite cores and found results simi-
to an optimum flux u = 5.09 × 10–4 m/s corresponding to an lar to those shown by Fig. 16-14. The results for
injection rate of 0.8 bbl/min. This rate is usually sustainable dolomite represent a striking confirmation of predic-
without fracturing the matrix.
Thus, most calcite formations are treated using rates near tions based on Eq. 16-39.
the maximum that the matrix will accept, except perhaps in The field implication is that acidizing in shallow
cool, shallow formations. Once the treatment has been initi-
ated, it may be beneficial to increase the rate to propagate dolomite formations should be conducted at low
the wormholes created by the initial acid contact. rates. High rates result in a multiple wormhole pat-
tern that does not penetrate far into the formation and
appears as uniform acid invasion dissolving the face
• Formation composition of the wellbore, which is inefficient for removal of
The reaction rate of HCl with dolomite is much skin effect damage compared with producing a few
slower than that with calcite (see Table 16-4). This dominant wormholes that penetrate into the forma-
being the case, Eq. 16-39 indicates that unless the tion. In deeper dolomite formations, the rate may be
acid flux is greatly reduced, many of the native pores increased to some extent because the reaction tem-
are likely to exceed AT in size and be candidates for perature increases with depth. The increased opti-
wormhole initiation. Thus, closely spaced multiple mum rate with increasing temperature shown in
wormholes are likely to form, producing a highly Fig. 16-14 confirms predictions based on Eq. 16-39.

16-26 Fundamentals of Acid Stimulation


greater the rate of acid diffusion toward the wall, the
80
Room temperature
lower the concentration of acid at the tip of the worm-
125°F hole and the slower its rate of propagation. Increased
70 170°F
fluid-loss rates serve to convect acid to the wall and
Pore volumes to breakthrough

Dolomite
1-in. diameter × 5-in. long at the same time reduce the acid flux reaching the tip,
60 3.4% HCl (1N) thereby decreasing the rate of propagation. Taking into
account both fluid loss and diffusion, Hung (1987)
50
found that for a constant injection rate, the rate of
extension of a wormhole begins to decrease as the
40 wormhole length increases. The length appears to ulti-
mately reach a plateau, as shown by Fig. 16-15 (Hung,
30 1987) but never actually ceases to grow. Hung attrib-
uted the diminishing growth rate entirely to fluid loss.
20 Thus, it is anticipated that wormhole penetration will
0 50 100 150 200 250 300 350
essentially cease after a certain length has been
Injection rate (mL/hr)
attained as long as the injection rate is fixed. In long-
Figure 16-14. Effect of temperature on the variation of vol- core experiments, an ultimate length was observed by
umes to breakthrough with the injection rate for dolomite. Bazin et al. (1995).
Hung calculated that the wormhole evolves in shape
depending on the local rate of acid reaction and fluid
loss and the rate of fluid injection. Once the wormhole
16-4.5. Propagation of wormholes length stabilizes, the acid that is injected serves primar-
Once wormholes are initiated in the rock surrounding ily to increase the diameter. Because Hung’s model
the face of the wellbore or perforation, it is desirable to does not account for the meandering nature of worm-
extend them into the formation as far as possible with holes caused by small-scale heterogeneities in the rock
a given volume of acid. The skin effect should be or the creation of side branches, it tends to overpredict
reduced within the regions penetrated by wormholes. wormhole length.
To promote understanding of the factors governing the Daccord et al. (1989) recognized the importance of
rate of extension of a wormhole, Hung et al. (1989) propagating the wormhole to the fullest extent possible
modeled wormhole growth by considering it to be a and proposed a model based on laboratory experiments
cylinder with fluid loss as depicted by Fig. 16-10.
Hung et al. took into account a number of factors,
including the contributions of both acid diffusion and 125
Injection rate
convection resulting from fluid loss to the walls of the 0.005 cc/s
0.003 cc/s
wormhole where the acid reacts. These are important 0.001 cc/s
100
factors because the acid reactions in a wormhole are,
in general, limited by mass transfer as contrasted to
those in natural pores, which are controlled by the 75
Length (cm)

reaction rate.
The rate of wormhole extension is determined by
50
the amount of the acid arriving at the tip:
dL uCβ uC
= e e 100 = e e Ac , (16-40)
dt (1 − φ)ρrock φC0
25

where the subscript e refers to conditions evaluated at


0
the end or tip of the wormhole, ρrock is the density of 0 80 160 240 320 400
the rock, and L is the length of the wormhole. This Time (s)
equation shows the importance of diffusion, acid con-
vection and fluid loss on wormhole propagation. The Figure 16-15. Predicted wormhole length (Hung, 1987).

Reservoir Stimulation 16-27


that differs from that proposed by Hung. Daccord et does not indicate a plateau value as the wormhole
al.’s expression for the rate of wormhole propagation lengthens. Thus, the equation is applicable to short
in linear systems is wormholes where fluid loss is not a factor, but it
2/3 should not be used for the prediction of wormhole
dL aAc  q 
= , (16-41) penetration length.
dt Aφ  D  Thus, none of the existing models for the rate of
where a is a constant determined experimentally, D is wormhole propagation is strictly correct. Because
the molecular diffusion coefficient, and A is the cross- wormhole length is thought to be a crucial factor in
sectional area of the wormhole. Daccord et al.’s model determining stimulation, better models incorporating
considers the influence of acid diffusion but does not the important features of the ones that have been pro-
take into account fluid loss; therefore, this equation posed are required.

16-28 Fundamentals of Acid Stimulation

Potrebbero piacerti anche