Vous êtes sur la page 1sur 17

Scale Prediction for Oil and Gas Production

Amy T. Kan and Mason B. Tomson, Rice University

Summary
Scale prevention is important to ensure continuous production
from existing reserves that produce brine. Wells could be abandoned prematurely because of poor management of scale and
corrosion. The objective of this paper is to present an overview
of scale prediction and control and the current research at Rice
University to solve these problems. In this paper, the challenges
of scale prediction at high temperature, high pressure, and high
total dissolved solids (TDS) and an accurate model to predict pH,
scale indices, density, and inhibitor needs at these conditions are
discussed and reviewed: specifically discussed are (1) the various
scale types found in oil and gas production and the condition under
which they form; (2) the relationship of pH, alkalinity, organic
acids, carbonates, and CO2 distribution; (3) the temperature (T),
pressure (P), TDS dependence of the thermodynamic equilibrium
constants and activity coefficients; and (4) the accuracy of the
Pitzer ion-interaction model-based scale-prediction algorithms and
their application. On the basis of a simple propagation of error
estimation, the overall estimated error for calcite saturation index
(SI) is 0.1. This algorithm has been validated with literature
solubility data for six minerals in the T, P, and TDS range of 0
to 200C, 0 to 15,000 psia, and 0 to 350 000 mg/L TDS; for pH
data at 25 and 60C; and density of weighting fluids with density
between 8 and 12.7 lbm/gal.
Introduction
By all accounts, oil and natural gas will be the major components of global energy production for decades to come. Although
proved oil and gas reserves may be declining, oil and gas production can increase with the development of new technologies to
extract energy from unproved reserves and improved economics
of production from both known reservoirs and marginal fields.
For example, oil and gas have gone to deeper and tighter formations with the advance of new production technologies. These
new developments in oil and gas production also bring challenges
in scale prediction and inhibition [e.g., at high T (150 to 200C),
P (1,000 to 1,500 bar), and TDS (greater than 300 000 mg/L)
commonly experienced at these depths]. Brines from different
zones are often mixed in the production tubing, subsea tieback,
or a common facility. Often the oil, gas, and brine composition in
each zone is quite different and consequently will cause significant
scaling. Steam, supercritical carbon dioxide, seawater, or alkaline
surfactant flooding are employed to mobilize residual or heavy
oil. Many of these cost-saving measures can result in significant
scaling problems. Scale prediction, control, and treatment are vital
to the success of these processes.
A major component of scale and corrosion management is the
ability to accurately predict the brine chemistry, pH, and scaling
tendency of a production system. The typical variables in oil and
gas fluid components are shown in Table 1. Because of the complexity of the system, scale prediction is not as straightforward
as one might imagine. In this paper, we attempted to address the
following: (1) the theoretical background of scale prediction and
reliability; (2) the impact of brine salinities, composition, and Ts
on pH, scale, and corrosion; (3) the impact of organic acids; (4)
the reliability of scale prediction at high T and P, and in complex
brine; (5) the impact of hydrate inhibitors on scale formation and
Copyright 2012 Society of Petroleum Engineers
This paper (SPE 132237) was accepted for presentation at the CPS/SPE International Oil &
Gas and Exhibition, Beijing, China, 810 June 2010, and revised for publication. Original
manuscript received 19 March 2010. Revised manuscript received 26 March 2011. Paper
peer approved 10 May 2011.

362

prediction; and (6) the prediction of scale inhibitor needed. Experimental data to validate the model prediction of pH, SI, and density
calculations are also examined, and an estimate of error because of
analysis is determined for a scale-prediction algorithm on the basis
of Pitzer theory of ion interaction to illustrate the interrelationship
of these parameters.
Common Scale Types Found in Oil and
Gas Production
Common oilfield scales can be classified into pH-independent
and pH-sensitive scales. The scaling tendency of sulfates (calcium sulfates, barite, and celestite) and halite scales are not a
strong function of brine pH. Although sulfate scale formation is
not pH sensitive, often the inhibitor reaction is sensitive to pH, as
will be discussed. The carbonates (calcite, dolomite, and siderite)
and sulfide scales are acid soluble, and their scaling tendencies are
strongly influenced by the brine pH. For pH-sensitive scales, the
scale prediction is more complicated because issues that control the
brine pH also affect their scaling tendencies. It should be noted that
at constant T, the solubility of all mineral scales decreases when
P decreases, more so for divalent ions.
Barite. Barite is the mineral name for barium sulfate (BaSO4).
When barite is formed rapidly (relative to geologic times)
from produced water it often contains up to one strontium ion
(Sr2+) per approximately seven barium ions (i.e., approximately
Ba0.88Sr0.12SO4). Sometimes, radioactive radium ions (226Ra2+) also
substitute into the barite crystal lattice and are the major source of
naturally occurring radioactive materials (NORMs).
Barite scale formation can occur because of a pressure drop during production from a reservoir wherein the brine is saturated with
BaSO4, but barite scale formation is generally a consequence of mixing high-sulfate waters such as seawater with formation water during
waterflooding operations or of mixing brine from a high-barium
zone with brine from a high-sulfate zone. Essentially, all minerals
in a reservoir are at equilibrium with the undisturbed connate, or
natural, water. The US Geological Survey (USGS) performed a
survey of log (barium) vs. log (sulfate) for a large group of wells
(Kharaka et al. 1988). The results showed that these wells produce
brine that is saturated with respect to barite at downhole conditions,
within experimental error. Often, either Ba2+ or SO42 concentration
is very low (the limiting reagent effect). Low concentrations of
Ba2+ (<1.0 mg/L) or SO42 (<15 mg/L) are difficult to measure. This
probably accounts for some of the scatter in field data. Therefore,
it is suggested that when production takes place from a single zone
and the well is not flooded, the brine is probably at saturation with
barite in the reservoir. The SI of barite increases as both the T and
P decrease, but it can be demonstrated that the typical T and P drop
from a well producing from a single zone would rarely be sufficient
to cause barite scale to nucleate and grow in the tubing.
Generally, barite scale occurs when either brines from multiple
zones or brines from different wells are mixed, or when there is
flooding of high-sulfate water, as occurs with seawater flooding.
When multiple zones are produced and it turns out that one of the
brines is high in sulfate and the other is high in barium and they
mix at the well bottom, barite scale forms at the bottom of the well,
but is generally delayed for a few feet until mixing is complete.
In both of these cases, the SI value for barite can be exceptionally high and large amounts of scale can form quickly. This can
even produce a supersaturation level that is not controllable with
inhibitors, regardless of the concentration of inhibitor. These are
probably the most dangerous conditions, but unfortunately they
appear to occur rather commonly.
June 2012 SPE Journal

TABLE 1RANGE OF VARIABLES TO BE CONSIDERED IN OILFIELD SCALE PREDICTION AND CONTROL


Temperature

Pressure

Oil

Brine

25 to 500F
or
4 to 260C

14 to 30,000 psi
or
1 to 2,000 atm

Aromatics
Paraffins
Waxes
Asphaltenes
Naphthenic acids

Na , K , Ca ,
2+
2+
2+
Mg , Ba , Sr ,
2+
2+
2+
2Fe , Zn , Pb , Cl , SO4 ,
HCO3 , Br ,
HS , HAc (acids),
SiO2
MeOH,
MEGs,
0.00 to 400,000 mg/L TDS

Flooding produces a very special and rather peculiar scenario


for barite scale formation. Because the injected brine flows over
the formation solids for considerable distances (excess of a few
feet is sufficient), the injected flowlines will establish equilibrium
with barite in the flow field. Therefore, the special problem of
barite scale in flooding is just a transient problem that occurs as the
produced brine goes from mostly connate brine to mostly injected
brine. Once the well has experienced injection-brine breakthrough,
the scale problem likely disappears, or simply is the same as if the
brine were saturated at downhole conditions. The important issues
of concern are generally related to the few feet around the injection well and the production well where the conditions can change
rapidly (Shen et al. 2009). In subsea completions, often methanol
or monoethylene glycol (MEG) is injected to prevent hydrates,
and these hydrate inhibitors can very easily cause the barite scale
problem to be very severe.
Three precipitation regimes can be identified in a well, depending upon the SI of the produced brine (Fig. 1). The exact critical SI
values that divide these three regions are highly dependent upon the
temperature. At lower temperatures, the critical SI is much higher
as indicated semiquantitatively in Fig. 1.
A. If barite SI in the perforation region at the bottom of the
well is greater than approximately 1.5 (120C), then there will be
virtually instantaneous precipitation in the mixing brines. This will
reduce the SI value, but with a large mass of crystals produced in
the process. Then seeded growth precipitation will commence in

SI > +1.5 (120 C)


SI > +3 (25 C)

Processes
+

Equil, rate, and masstransfer processes:


gas, solids,
redox, corrosion,
bio., minerals,
NORMs
Inhibition kinetics and
concentrations for eight
different inhibitors and
blends

the solution and on the tubing walls at the bottom on the well. This
condition can be difficult to inhibit with chemical intervention.
B. If barite SI in the perforation region at the bottom of the well
is greater than approximately 0.3 to 1.5, then, in the absence of
inhibitors, there will be nucleation and precipitation on the tubing
walls at the bottom of the well. The kinetics of this scale growth on
the tubing walls will be quite rapid, and scale is expected to form
and build up at the bottom of the tubing, as often occurs. Because
of low barite solubility, the mass of barite scale formed per unit
brine is typically small and may not be detected for a long time.
For this reason the pattern of mineral buildup is often such that
by the time the brine reaches the surface, it is calculated to be at
equilibrium, and one is tempted to suspect that there is no problem.
In these cases, the buildup of barite can sometimes be calculated
by taking the reservoir brine component values minus the surface
values and summing for the barrels of brine produced.
C. If barite SI in the perforation region at the bottom of the well
is less than approximately 0.3, then nucleation will be delayed up
the well. This also applies to wells wherein the brine is at equilibrium with barite in the formation, as can occur in the absence
of seawater flooding. As the brine flows up the tubing, both lower
pressure and lower temperature increase the SI value for barite. All
else being equal, the rate of nucleation and scale growth increase
as the SI value increases. However, as the temperature decreases,
barite precipitation rate decreases quite rapidly. This can cause
barite to nucleate and grow farther up the tubing toward the sur-

Above some critical SI value nucleation and


crystal growth will be spontaneous and can not be
prevented by added SIs. This critical
SI value for barite is not well known and depends
upon T, pH, and the inhibitor selected.

SI + 1.5 (120 C)
SI + 3.0 (25 C)

Solutions in this region can be inhibited


with barite SIs. Nucleation and
crystal growth can be delayed for long
periods of time, minutes to days or longer.

SI + 0.3 (120 C)
SI + 1.3 (25 C)
SI = 0.00

2+

Precipitation and
growth occur only on
seed crystals of barite.
Equilibrium

If the SI < 0, only dissolution can occur.


SI = negative
Time (Not to Scale)
Fig. 1Schematic to illustrate the relationship of SI and kinetics. The lines are time to precipitate or dissolve to equilibrium for
barite. A similar diagram could be constructed for any type of scale formation, except the specific T and SI values would change.
Typically, the more soluble a mineral, the lower the corresponding SI values.
June 2012 SPE Journal

363

face or this reduced temperature can prevent nucleation and scale


buildup altogether. Scale forms either toward the top of the tubing
or not at all.
Calcium Sulfates. Three calcium sulfate phases commonly occur
as scales. They differ by the number of water molecules in the
crystal formula. Gypsum (CaSO42H2O) is the stable calcium
sulfate phase at low temperatures (from room temperature up to
approximately 40 to 90C). Above approximately 120C, anhydrite (CaSO4) is the commonly reported phase. At intermediate
temperatures, hemihydrate (CaSO4H2O), also called plaster
of Paris, is often reported in brines with high TDS content. The
transition temperatures are strongly dependent on the salt content
of the brine and other specic chemical species. At any specic
temperature, a higher concentration of salt will tend to favor the
formation of a solid calcium sulfate phase with fewer water-ofhydration molecules in the crystal formula. The details of how
and which solution conditions affect the phase relationships for
the calcium sulfates are not known well enough to be predictive,
but it is certainly related to the water activity. In the presence of
methanol or MEG, as hydrate inhibitors, anhydrite was observed
at much lower temperatures than expected because of the lowering water activity in the presence of methanol or MEG (Kan et al.
2003; Lu et al. 2010).
The precipitation of calcium sulfate scale phases is often a consequence of mixing of incompatible waters during waterflooding.
Also, calcium sulfate precipitation can result from a pressure drop
when production is from a reservoir where the brine is saturated
with calcium sulfate, or from an increase in temperature during
the processing of the brine on the surface (e.g., heater), during
membrane filtration, during steamflood, or when large quantities
of thermodynamic hydrate inhibitors are used for hydrate control
(Kan et al. 2002; Kan et al. 2003; Lu et al. 2010).
Celestite. Celestite is the mineral name for SrSO4. The solubility of celestite salt is less studied than calcium sulfate or barium
sulfate (Kaasa 1998). There is a wide variability in literaturereported celestite solubility (Collins and Davis 1971; Reardon
and Armstrong 1987; Felmy et al. 1990; Howell et al. 1992). Two
issues appear to contribute to the discrepancy: the presence of
smallest-sized particles that control the solubility and the susceptibility of this mineral to surface poisoning. Monnin and Galinier
(1988) reviewed the solubility of celestite in complex solution
matrices. They concluded that the celestite solubility in NaCl is
well behaved according to the Pitzer model, but its solubility in
sulfate brine is not. The unexpected solubility is attributed to the
effect of strong Sr2+/SO42 ion-pair interaction. We are currently
testing these ideas.
Halite. Halite is the mineral name for NaCl. The Pitzer theory was
derived from the NaCl thermodynamic data measured by Pitzer. The
scale prediction of halite should be accurate up to 300C. Halite
solubility is 6.275 molal or 322.6 g/L at 122F or 50C (Pitzer et al.
1984). Therefore, halite does not typically form in less than approximately 320 000 mg/L TDS brine in oileld conditions. When the
produced brine is near saturation with halite downhole, a small drop
in temperature can cause halite to precipitate. Therefore, halite scale
can be problematic for a gas well with high TDS brine and cooling
as the pressure drops. Because of the high halite solubility, the mass
of NaCl formed is large if precipitation occurs. Wells tend to be
plugged fast when halite SI exceeds 0.0. The solubility of halite in
pure water and in electrolytes has been studied thoroughly (Pitzer
et al. 1984). Note that halite solubility is reduced in the presence of
methanol. Halite scale can become a problem at much lower TDS
values, approximately equal to and greater than 180 000 mg/L in
the presence of methanol (Kan et al. 2003).
Silica. Silica and silicate chemistry is quite complicated and highly
varied in rocks and reservoir brines. Silicate scale is rarely a problem in lower-temperature production. How low is low is not
completely clear. As long as the temperature is below 300 to 400F,
it is unusual to have silicate scale formation. Because silicates
364

typically have a rather simple monotonic increase in solubility with


temperature, the reverse is also true (i.e., nearly all produced waters
are supersaturated with silicates as temperature decreases). Yet, we
rarely see silicate precipitates. The silica monomer species in solution tend to be slowly polymerized to larger and larger polymers
with time, but this reaction is much slower at lower temperatures.
The temperature of transition to silica problems seems to be near
400 to 500F (204 to 260C).
Sulfides. The sulde scales include FeS, FeS2, PbS, CaS, and ZnS,
to mention a few. Sour elds, which produce a substantial amount
of hydrogen sulde (H2S) gas, tend to precipitate sulde minerals.
Many of the deeper and hotter gas elds contain 25% by volume
or more H2S. H2S is very corrosive to most steels and tends to
form metal sulde precipitates. Several investigators have reported
the formation of zinc sulde (ZnS, or sphalerite) at depth, even in
sweet, or low-sulde, brines. Even though discrete formulas such
as ZnS and FeS are used in discussing these suldes, it should
be emphasized that these are only idealizations and that in most
cases the actual solid compositions are not well known. Further,
little is known about what controls the conversion of acid-soluble
FeS to acid-insoluble FeS2 (pyritization). Considering the importance of these materials to production, surprisingly little is known
about the prediction and control of metal sulde deposits. The
research groups of the present authors and others (Okocha et al.
2008; Sun et al. 2008; Okocha and Sorbie 2010) are developing
the fundamental knowledge to better understand issues related to
sulde scale control.
Siderite. A special relationship has been reported to exist between
iron carbonate (FeCO3) and iron oxide formation and corrosion
control as a function of temperature (Sontheimer et al. 1981; Dunlop et al. 1985; Tomson and Johnson 1991). At low temperatures,
FeCO3 is slow to form, but so is corrosion. As the temperature
increases to approximately 100C, the rate of FeCO3 formation
accelerates. Somewhere between 90 and 110C, FeCO3 will begin
to decompose and form magnetite, Fe3O4, which is a better corrosion-prevention lm. At higher temperatures, the corrosion rates
tend to decrease to some lower limiting value. The relationship of
calcite scale formation to corrosion control has already been mentioned, but Sontheimer et al. (1981) suggested that the important
rst phase to precipitate is generally FeCO3, not CaCO3 (Tomson
and Johnson 1991; Greenberg and Tomson 1992). It is generally
observed that in the absence of corrosion, the ratio of Ca2+ to Fe2+
is approximately 100:1, because of the corresponding Ksp ratios,
but essentially pure FeCO3 has been reported in several elds
around the world.
Many reports on the relationship of scale deposits to corrosion
control have been published, including the formation of metal
oxides, phosphates, phosphonates, chromates, and molybdates,
to mention a few (Ikeda et al. 1985; Burke and Hausler 1985).
There are numerous opportunities in the oil and gas business for
using such precipitates in corrosion-control programs, especially in
fields that produce substantial volumes (approximately 500 B/D, or
more) of brine. It is expected that more relationships between scale
and corrosion will be developed as the need for high-efficiency
corrosion control increases along with saltwater production.
Dolomite [CaMg(CO3)2]. To our knowledge, there is no way to
precipitate dolomite in the laboratory. It forms over thousands to
millions of years in the eld. To date, no one has been able to precipitate dolomite in the laboratory at any SI. The SI is more related
to dissolution of dolomite to equilibrium and indicates whether a
formation is probably saturated with dolomite.
Calcite. Calcite, essentially CaCO3, is the most common type of
scale found. The crust of the Earth contains approximately 12%
by weight of calcite and closely related carbonates. If the pressure drop from bottomhole to the surface is a factor of 5 to 10, or
greater, it is reasonable to expect that calcite scale will form; this
is particularly true for sweet wells that are not being waterooded.
Calcite scale formation is generally a consequence of the pressure
June 2012 SPE Journal

pH of Pure Water, No Salt

7.5
7
6.5
6
5.5
0

25

50

75

100 125 150 175 200 225


T (C)

Fig. 2Plot of pure-water pH vs. temperature.

drop that accompanies production. At pressure below the bubblepoint, this pressure drop removes carbon dioxide (CO2) from solution and increases the solution pH and causes calcite to precipitate.
Also, there is a secondary consequence of the pressure drop; the
inherent solubility of calcite in saltwater decreases as the pressure
decreases. Both of these effects tend to cause calcite to precipitate during production. Calcite crystals are composed mostly of
calcium carbonate (CaCO3), but often contain up to 20% iron or
magnesium carbonate, so that the average formula for calcite might
be represented as Ca0.8 to 1.0(Fe,Mg)0.2 to 0.0CO3.
Although naturally occurring calcite such as Iceland spar is
often essentially pure CaCO3, the scale formed from flowing brine
generally contains several mol% of iron. This coprecipitated iron is
often from corrosion products deeper in the well, but can also be a
result of naturally occurring siderite, FeCO3, or other materials in
the reservoir. Most divalent metal ions, such as manganese (Mn2+),
lead (Pb2+), and strontium (Sr2+), can substitute into the calcite
crystal lattice (Deer et al. 1983). Alsaiari et al. (2010) studied the
coprecipitation of iron and calcium carbonate. Calcium ions have a
strong influence in increasing the solubility of iron carbonate. On
the other hand, ferrous iron does not significantly affect the solubility of calcium carbonate. Although calcium carbonate has higher
solubility and also the SI of calcium carbonate was lower than that
of ferrous iron carbonate, calcium carbonate is the preferred phase
to precipitate and the precipitate has a higher molar ratio of Ca/Fe
than that in solution. Additional discussion on scale formation and
inhibition can be found in Frenier and Ziauddin (2008).
pH Definition and the Interpretation of
pH Value of Brine
pH determines the scaling tendency of the pH-sensitive scales
and is of value in interpreting other brine chemistry, such as in
scale, corrosion, or emulsion formation. Yet, the procedures used
to calculate or to measure pH are often of questionable reliability
in oilfield applications. The pH is defined to be equal to the negative of the logarithm (base 10) of the hydrogen-ion activity on a
molality scale (moles of H+ per 1.000 kg of water) with the reference-state activity coefficient equal to 1.000 at infinite dilution in
water [i.e., pH log10 (aH + ) = log10 (mH +  H + )].
There are at least three fundamental problems with the concept
of pH and they can only be approximately addressed thermodynamically (Bates 1973): (1) It is impossible to separate the required
activity coefficient product ( H +  Cl ) ; (2) there will always be a
junction potential Ej(V) between the solution and the reference
electrode; and (3) the infinite-dilution transfer potential from one
solution to another (e.g., pure water to methanol/water) cannot be
determined. With that realization, a practical approach is adapted to
standardize the pH measurement by comparing the electromotive
force (emf) of a sample to the emf of a standard reference buffer
measured with a glass electrode under well-defined conditions. We
use the operational definition of the pH as is used by the National
Institute of Standards and Technology, which is pH PH(S) +
(EX ES)/(2.303RT/F), where pH is the value read on a pH-meter,
pH(S) is the value assigned to as standard reference buffer [e.g.,
pH(S) = 4.008 for 0.05 m potassium hydrogen phthalate buffer at
June 2012 SPE Journal

25C], ES is the voltage in the standard reference buffer, and EX


is voltage in the unknown solution. The term (2.303RT/F) is the
Nernst factor, where F is the Faraday constant (= 96,484.6 coulombs/equivalent), R is the gas constant [= 8.31441 J(molK)1],
and T = temperature in K. Then, (2.303RT)/F = 0.05916 V at
298.15K. In short, pH is the value read on a pH-meter, when
calibrated as in the preceding. How consistent are the National
Bureau of Standards within themselves and how closely do they
correspond to the thermodynamic concept of pH? The uncertainty
of pH(S) appears to be approximately 0.006 at 25C and somewhat larger at other temperatures. The uncertainty relative to the
thermodynamic standard appears to be 0.003 pH units.
Because pH is determined on a relative scale to a standard
reference buffer, the pH measurement must be made under welldefined conditions in the field. Furthermore, any oxidation and
precipitation reactions of the brine after it is collected can alter
both the total alkalinity of the brine and reduce the pH. We have
proposed a sampling procedure by first lowering the temperature
of the brine with ice water while brine is still pressurized and
then dropping the pressure of the brine under a blanket of CO2
gas. If a sample is stored refrigerated, this will yield a chemically
unaltered brine at full strength (Tomson and Oddo 1996). Given
the difficulties in obtaining meaningful pH values for field brine
samples, we have developed rigorously accurate equations to eliminate the need to use measured pH in SI calculations. Essentially,
we mathematically substitute the measured PCO2 for pH, as will be
discussed later. To be exact, if pH is the chosen parameter for scale
prediction, the conditions at which pH is measured must be known.
We recommend that pH be measured at standard temperature and
pressure (STP) and at equilibrium with gas of composition exactly
the same as the produced gas. This is approximately what would
be measured at a low-pressure separator on the surface after the
gas has cooled to room temperature and the pressure is reduced.
Next, the measured pH should be corrected for both the salinity
and cosolvent effects on the electrode junction potential in order
to estimate the true pH, as needed for calculations.
For any real solution and pH measurement there is always a
junction potential, Ej, because of the difference in the salt content of the solution and the reference electrode filling solution
{i.e., [ E (V ) = E o ( RT F ) ln(mH + mCl  H +  Cl ) + E j ]}. The junction
potential Ej is a result of two processes. First, there is generally
a concentration gradient from the solution to the reference electrode filling solution, at the liquid junction, and this concentration
gradient causes a voltage difference. Second, the cations and
anions typically diffuse at different rates, given by their transference numbers or essentially their mobility (/F). KCl is used
as a filling solution because the mobility of K+ and Cl are very
similar [73.52(K+) vs. 76.34(Cl) Scm2/eq.] Values of Ej can be
estimated by use of the Henderson equation (Bates 1973; Bard
and Faulkner 1980). Also, Ej can be measured using HCl, and
the values generally vary from 0.02 to 0.5 pH units, depending
upon ionic strength. For dilute aqueous solutions, Ej is generally
approximately constant, 1 mV, or less, relative to the standard
buffer solutions. The authors established a correlation equation to
estimate the change in junction potential as a function of salt concentration (Kan et al. 2002; Lu et al. 2010), pHmeter reading = pHtrue
pHj = pHtrue 0.13(I)1/2, where pHj is the change in junction
potential in unit of pH and I is the ionic strength in molality. If the
brine contains hydrate inhibitors (e.g., methanol, MEG) a further
complication arises related to the precise meaning of the reference
state for pH. A detailed discussion on the cosolvent effects has been
given previously (Kan et al. 2002; Lu et al. 2010).
Effect of Temperature, Pressure, Gas-Phase
CO2 and H2S, and TDS on Brine pH
The pH of water can be affected by T, P, CO2 and H2S contents,
and TDS. The following calculation is used to illustrate the effect
of T, P, and CO2 in the gas phase, and TDS on brine pH. The
pH of pure water is equal to the negative logarithm of the square
root of the ionization constant of water [i.e., pH = log10(Kw1/2)],
and at 25C Kw = 1014 and pH = 7.00. Fig. 2 is the calculated
365

pH of pure water vs. T with no added salt, which is the predominant part of the T dependence of the pH. The salt effect on
pH can be derived from solving the equilibrium (K w = aH + aOH )
and the charge-balance ([H+] = [OH] equations simultaneously,
pH = log10 (aH + ) = log10 ( K w1/ 2 ( H +  OH )1 2 ). That is, the pH
differs from the pure-water value, just given, only by the square
root of the ratio of the activity coefficients of the hydrogen and the
hydroxide ions. The activity coefficient of the hydroxide ion is almost
always less than that of the hydrogen ion. At 1.0 M NaCl, or 23 000
mg/L Na+ and 35 450 mg/L Cl, and 25C: H+ = 0.876 and OH =
0.610, and pH = log10[10140.876/0.610) = 6.92. Because of the
salt effect on pH electrode junction potential, the pH reading will
be that corrected [i.e., pHmeter reading = 6.92 0.13(1.0 M)1/2 = 6.92
0.13 = 6.79]. In the presence of CO2, the pH will decrease and the
effects of T and P can become large and cause the steel to corrode
(i.e., the classical CO2 corrosion). In the presence of 1% CO2 in the
gas and at room temperature and 1-atm pressure, the pHmeter reading =
4.68. Increasing the temperature to 200F but remaining at 1 atm
pressure raises the pH from 4.68 to 4.96. This is because of the fact
that at the same PCO2 = 0.01 atm, the amount of CO2 that is pushed
into solution is less than that at 77F (25C). There is a secondary
impact of the change in the ionization constants of carbonic acid and
of water, but the primary effect is because of the decreased solubility of CO2 in water at the higher temperature. Next, raising the P
to 10,000 psi at the same T of 200F would lower the pH to 3.76.
This is because the increased pressure greatly increases the PCO2 and
the amount of CO2 that is pushed into the salt water. Overall, these
calculations illustrate the effect of temperature, pressure, TDS, and
CO2 on pH. Water at neutral pH can become corrosive downhole
or scale can form at the surface from nonscaling calcite-saturated
brine. A number of operators developed downhole pH sensors to
measure reservoir pH directly (Breng et al. 2003; Raghuraman
et al. 2007). However, the accuracy of such methods has not been
well documented and is difficult to validate, and international
accepted standards have not been adopted.
Relationship of Alkalinity, Bicarbonate,
Sulfide, and Organic Acids
Bicarbonate is a crucial variable in carbonate-scale prediction,
and it buffers the brine pH. Because bicarbonate concentration
can change with pressure down a well, it is not a conservative
parameter to be used in scale prediction. The change with pressure
is strongly dependent on the specific PCO2, the relative volume of
water, gas, and oil produced; and the total carboxylic acids present.
A useful way to address this problem is by using the concept of
alkalinity (Alk). The advantage of this concept derives from the
fact that although the actual concentration of the weak-acid anions
and pH change with pressure and PCO2, the Alk is constant, as long
as no net acids or bases are added such as CaCO3 precipitates or
dissolves. Note that neither adding nor losing electronically neutral
CO2, H2S, or SiO2 causes any change in Alk. Once the total Alk
and organic acids concentrations are determined by the procedure
discussed, the theoretically corrected bicarbonate concentration
and/or pH at any given operating temperature and pressure can
be calculated.
The origin of Alk is reactions of weak acids from diagenesis
and metabolism with minerals in the formation, such as CaCO3,
Fe(OH)2, MgSiO3, and Na2SiO3, to form weak-acid buffer solutions,
as shown by the following reaction scheme:
minerals
(H 2 CO3 + HAc + H 2S + HCl)aq metal

(HCO3 + CO32 + Ac + HS + S2 )
This is essentially a statement of the charge balance for a solution (Alk =Na+ + K+ + 2Ca2+ + 2Mg2+ Cl 2SO42 ) (i.e.,
the net alkalinity is the sum of strong-base cations minus the sum
of strong-acid anions). Alk is often defined as the total capacity
of the solution to neutralize strong acids. For a system containing
only carbonate, sulfide, and carboxylic acid species, Alk is defined
by the following expression:
366

Alk [HCO3] + 2[CO32] + [HS] + 2[S2]


+ [Ac] + [OH] [H+]. . . . . . . . . . . . . . . . . . . . . . (1)
The twos in front of [CO32] and [S2] express the fact that one
mole of each ion is capable of neutralizing two moles of acid,
H+. The term [Ac], acetate ion concentration, is the sum of all
the concentrations of ionized carboxylic acids in the brine (e.g.,
{[Acetate] + [Propionate] + [Butyrate] + }). If the brine pH
is above approximately 10 to 11 pH (which is very unusual, but
possible), then it is necessary to add contributions to Alk from
boric acid (H3BO3), silicic acid (H4SiO4), and ammonia (NH3).
For the effect of phosphates and polymers to be significant, the
concentrations would have to be more than 10 to 100 mg/L,
active. There is considerable debate about the numerical value of
the second ionization constant of H2S; values range from K2,H2S =
1013 to 1018. In either case, the concentration of S2 is essentially
never of importance in the calculation of Alk. For illustration
purposes, all these minor components are omitted in Eq. 1 and in
further discussion.
The predominant Alk term is often [HCO3]; this is especially
true in the absence of substantial H2S, organic acids and when the
pH is less than approximately 8 to 8.5 pH (the pH range wherein
CO32 becomes significant). However, oilfield brines often contain
aliphatic carboxylic acids, up to 5000 mg/L. The highest concentrations of carboxylates tend to be in waters from reservoirs
at temperatures of 80 to 100C (Carothers and Kharaka 1978;
Lundegard and Kharaka 1990; Lundegard and Land 1993). When
brine contains a substantial amount of organic acids, the conventional assumption that bicarbonate concentration is equal to Alk
is no longer valid. This is illustrated by the following example. At
200F, 1,000 psi, and 1 M TDS 60 000 mg/L; the conditional
equilibrium constants (He et al. 1997) are K1H 2 CO 3' = 10 5.94 , K HCO 2 ' =
'
102.04 atm/M, K HAc
= 104.48. In this specific example, the Alk is
essentially equal to the sum of bicarbonate and acetate, Alk =
[HCO3 ] + [AC ] = (10 7.98 PCO2 ) aH+ + THAc 10 4.48 (aH+ + 10 4.48 ) .
Because the Alk and the THAc are both constants and the PCO2
changes because of the flow of the well, the pH varies with the
PCO2. When the PCO2 is low, toward the surface of the well, the
pH will be higher or aH + will be lower and the second term on
the right-hand side will become THAc1.00 ( = [AC]) (i.e., all the
acetic acid will be ionized and [HCO3 ] will be less than it otherwise would be if the Ac concentration had been ignored). Then,
as the gas is compressed down the well, the CO2 in gas phase will
be forced back into the brine, which will lower the pH. At lower
pH, the ratio of the second term becomes smaller. At pH = 4.48,
only approximately half of the THAc will be ionized, [Ac] = THAc.
Therefore, the actual amount of [HCO3 ] will increase as the gas is
compressed down the well with a corresponding quadratic change
in SI(calcite) (see calcite SI in Eq. 7).
Fig. 3 plots the HCO3 concentration and brine pH as a function of PCO2 and acetic acid concentration in the brine. As shown
in Fig. 3b (top line), the bicarbonate concentration is essentially
unchanged between 0.01- to 100-psi PCO2 in the absence of organic
acids. However, the bicarbonate concentration changes considerably between 0.01- to 100-psi PCO2 when the organic acids are >
25% of total Alk. For example, with Alk = 610 mg/L = 0.01 equiv/l
and THAc = 0.009 M with 5% CO2 (STP), the pHSTP 6.00. At this
point, [HCO3 ] 0.00112 M and [Ac ] 0.00888 M. In the well at
100 times this pressure, PCO2, gas = 5 atm: [HCO3 ] 0.00370 M and
[Ac ] 0.00630 M at 4.85 pH. This shows that the [HCO3 ] has
increased by 330%. This will change the calcite SI by log(3.3)2 =
1.04 SI units vs. what would have been calculated without correcting for the acetic acid effect.
Simultaneous Alk and Weak-Acids Determination
Reliable methods to accurately measure true Alk are scarce, especially when organic weak acids are present that smear the titration
endpoint of bicarbonate. Oxidation and precipitation reaction of
the brine can also alter the true total Alk of the brine. Similarly,
measuring the organic-acid compositions require either a gas chromatograph or liquid chromatograph procedure, and the method is
June 2012 SPE Journal

TAc (M)

0.01
0.011

P-CO2, psia

0
10
0

10
0

20

0.
2

0.
05

0.
01

4.00

0.007
0.008

0.002

20

5.00

0.005
0.006

0.004

0.004
0.005
0.006
0.007
0.008
0.009

6.00

0.003
0.004

0.006

7.00

0.008

0.
2

pH

8.00

0.001
0.002

0.
05

9.00

0.01

0.
01

0
0.001
0.002
0.003

HCO3 conc., M

10.00

TAc (M)

0.012

P-CO2, psia

0.009
0.01
0.011

Fig. 3Plots of the variation of pH and HCO3 concentration vs. PCO2, gas and TAc concentrations, all at a final Alk = 0.01 eq/L 610
mg/L as HCO3. The equilibrium constants used are at 212F and 15 psia and 1 M I. Note that with any specific well, the change
in pressure will typically be by a factor of 100, or less.

tedious and expensive. We have developed a rigorous theoretical


solution of this problem and have proposed a corresponding experimental method to accurately measure both bicarbonate and weak
acids simultaneously, with minimal experimental difficulty and
no costly instrumentation (Kan et al. 2005; Tomson et al. 2006).
The new titration method is based on simultaneous analysis of the
titration curve determined at fixed PCO2 and emphasizes the titration shape (profile) instead of the endpoint inflection, as is done
presently using recommended procedures and apparatus (Fig. 4).
When the brine is purged with CO2 gas, each mole of HS will be
replaced by one mole of CO2 gas to form one mole of bicarbonate
with no net change in total alkalinity:
HS + CO2, gas HCO3 + H2S gas .
Typically, [CO32 ] and [OH ] concentrations are negligibly low.
The charge-balance equation of Alk titration can be expressed in
the following form after substituting the equilibrium constants and
activity coefficients:

Alk0 HCl( M ) =

K1H 2 CO 3 K HCO 2 PCO2 ,gas  g,CO2

 HCO

10 pH

TAc
10 pH  Ac

+ 1
K 
HAc HAc,aq

= a 10

pH

TAc
b 10 pH + 1

. . . . . . . . . . . . . . . . . . . . . . . . (2)
where Alk0 is the initial brine alkalinity (equiv.-liter1) and each
mole of added HCl exactly neutralizes one equivalent per liter of
alkalinity. Note that the activity coefficients and equilibrium constants in Eq. 2 are constant for a specific brine and, therefore, the
product of these constants in the first and second terms of Eq. 2 can
be regarded as two adjustable constants, a and b. The titration data
(HCl added vs. pH) can be fitted to Eq. 2 using a nonlinear leastsquares procedure by minimizing the sum of the squares of the
difference between the calculated and the experimental pH, subject
to the constraints of the function S min = in=1 ( pHcaic, i pH exp, i )2
being fitted to to determine the values (Alk0, TAc, a, and b). These
calculations can be performed using Microsoft Excel Solver utility. This method has been tested on three oilfield brines by four
independent laboratories of oil and service companies. Excellent
agreement was observed from the blind round-robin test results
(Table 2).
Typical reactions that may alter the initial brine Alk during
sample storage include oxidation and precipitation, as illustrated
in the following reactions:
June 2012 SPE Journal

Fe 2+ + 0.25O 2 + 2.5H 2 O Fe (OH)3,solid + 2H + (oxidation)


Ca 2+ + 2HCO3 CaCO3, solid + H 2 CO3, aq (precipitation)
Fe 2+ + 2HCO3 FeCO3, solid + H 2 CO3, aq (precipitation)
Fe 2+ + 2HS FeS + H 2Saq (precipitation)
In each of the reactions, two equivalents of Alk will be consumed per mole of reaction.
The following correction procedures are recommended for
brine that may be altered during storage.
1. Filter brine with a 0.45-mm filter.
2. Analyze total Alk and organic-acid concentration according
to recommended procedure (Fig. 4a).
3. Acidify and measure Fe and Ca content on the filter, and
calculate the equivalent Fe and Ca concentrations in the original
brine that form precipitates.
4. Alk in the original brine (mg/L) = Alktitration(mg/L) +
2.184[Feppt](mg/L) + 3.050[Cappt](mg/L).
A series of experiments was performed to test the feasibility of
this equation for correcting the effect of precipitation and Fe(II)
oxidation on Alk measurements. Fig. 5 plots the decline in measured Alk over time of a brine stored in two different sampling
bottles made of two different plastic materials [ plyvinyl chloride
(PVC) and high-density polyethylene (HDPE)]. The synthetic
brine contains NaCl (1 M), Fe(NH4)2(SO4)2 (154 mg/L as Fe2+),
and NaHCO3 (471 mg/L as HCO3). PVC bottles are significantly
better than HDPE bottles in terms of providing an oxygen barrier.
The oxygen permeabilities are 4 and 185 cm3-mil/100m2-24 hr-atm
for PVC and HDPE, respectively. Over a 10-day period, the Alk
declined by only approximately 8% when the brine was stored in
the PVC bottle with no headspace. The Alk declined approximately
35% when the same brine was stored in the HDPE bottle with no
headspace. The measured Alk in the HDPE bottle declined steadily
over time because of gradual Fe(II) oxidation and precipitation.
However, the loss of Alk because of oxidation and precipitation
can be corrected quantitatively. There is no statistical difference
among the corrected alkalinities and the total Alk initially present
in the brine, Alk0.
Finally, the various terminologies used in typical brine analysis
to express Alk (or ideas intended to express the true bicarbonate
concentration in a well) can also add to confusion. The total Alk
can be expressed as bicarbonate equivalents in unit of mg/L HCO3.
The following is a list relating the various terminologies to the
expression of bicarbonate equivalents in unit of mg/L HCO3.
1. If Alk or bicarbonate concentration is given in units of mg/L
of CaCO3, then total Alk (mg/L HCO3) = 1.22(mg/L CaCO3).
367

9
7
6

pH

6
Gas
Saturation
Bottle

Titration
apparatus

(a)

4
5
6
7
8

9/20/2000

C
Temp (F)
TDS (mg/L)

77.0
9,972

BJ

CO2 (%)

1.00

Std. Err.

Well Name

Total Alk
(mg/L)

986

Location

Carboxylic
acid (mg/L)

411

Field

r2

A
Sample ID
Date
Operator

0.9997

8.0
7.0
pH

6.0
5.0
4.0
3.0
0.00

0.01

(b)

0.01

0.02

0.02

HCl added (M, Exp)

"Exp"

"Calc"

Fig. 4(a) Schematic of the titration apparatus. The parts are (1) CO2/N2 gas cylinder, (2) gas regulator, (3) regulating valve, (4)
1/8-in.-outer-diameter Teflon tubing, (5) glass bottle with a GL-45 2 valve cap, (5) sparging filter, (6) 8-oz. jar with a 1.5-in. stirring
bar, (7) magnetic stirrer, (8) pH electrode, and (9) pH meter. (b) typical titration-results curve fitted by the BCC Total Alkalinity
and Carboxylic Acid Simultaneous Titration at Constant PCO2 program.

2. If both mg/L HCO3 and mg/L CO32 are given, then total Alk
(mg/L HCO3) =2.303(mg/L CO32) + (mg/L HCO3).
3. If you are given a value for phenolphthalein Alk as mg/L
CaCO3 in addition to bicarbonate alkalinity as mg/L CaCO3, then
total Alk (mg/L HCO3) = 1.22(phenolphthalein Alk + bicarbonate
Alk, mg/L CaCO3).
Relationship of pH, PCO2, Alk, and SI
The total moles of each gas component present in the gas, oil,
and brine can be used at all subsequent temperature and pressure
combinations to calculate pH as well as all other solution species at
the new temperatures and pressures. Only two out of the three key
parameters (i.e., PCO2, Alk, and pH) are needed to calculate the total
mass of CO2. Once the pH and PCO2 are measured or calculated,
the total moles of CO2 at specific volumes of gas, oil, and water
368

and at specific temperature, and pressure can be calculated from


Eqs. 3 and 4:

nT CO 2 = ngCO2 + noCO2 + nwCO2 + nwHCO3 + nwCO3


= PCO2 (Vg / ZRT + a2 ) , . . . . . . . . . . . . . . . . . . . . . (3)
2
CO 2
a2 =  CO 2,g ( K oCO
mw
/ g mo /  CO 2, o + K H

1
K H 2 CO 3 K1H 2 CO 3 K 2HCO 3 ,
+ 1
+

aH 2 CO 3
CO 2,aq aH HCO 3

. . . . . . . . . . . (4)

where Vg is the gas volume, Z is gas compressibility, and  is the


activity coefficient. Similarly, the total moles of H2S and hydrocarbon can be calculated. Using the principle of mass balance, the
June 2012 SPE Journal

TABLE 2ROUND-ROBIN TEST RESULTS OF THE RICE UNIVERSITY TOTAL ALKALINITY, CARBOXYLIC ACID SIMULTANEOUS

TITRATION AT CONSTANT PCO2 ANALYSIS


Organic
Acid
Std Brine Lab. Alkalinity Std
(mg/L) Dev.
#
#
(mg/L) Dev

Brine Lab. Alkalinity Std


#
#
(mg/L) Dev

Organic
Acid
(mg/L)

Std Brine Lab. Alkalinity Std


Dev.
#
#
(mg/L) Dev

Organic
Acid
(mg/L)

Std
Dev

1082

1234

14

401

557

47

514

10

15

12

1123

1254

13

386

418

50

484

32

49

33

1024

1138

10

377

15

229

64

529

91

60

73

1126

1175

10

386

494

60

532

107

41

89

1119

1172

12

377

483

31

525

35

1122

1041

46

329

486

134

543

1111

1220

24

337

381

34

516

1120

1270

34

353

432

39

516

1125

1248

354

431

60

1106

1235

GC analysis
Mean

1186
1133

297
222

0
0

1106

1199

18

359

435

58

520

32

25

27

Standard
error

10

22

31

10

15

13

Standard
deviation

32

69

13

24

93

31

17

43

24

35

By repeated substitution, it is possible to express Eq. 5 as


follows:

pH and PCO2, PH2S can be calculated from nTCO2 and nTH2S at other
temperatures and pressures.
The following illustration is to show how changes in volume
of water or gas will affect the calculated downhole pH by changing the relative fraction of gas-phase CO2 with pressure. At STP
conditions with 0.01-atm PCO2 and a gas/oil/water ratio of 2,000
scf/1 STB/10 STB, the solution pH is 7.25 and dissolved CO2
concentration ([CO2,aq]) is 2.8104 m. When the system is pressurized to 400 atm and is at 25C, PCO2 becomes 0.44 atm, more
CO2 dissolves into the aqueous phase ([CO2,aq]=0.0120 m), and the
solution becomes more acidic (pH 5.76). At the bubblepoint (513
atm) and 25C, the gas phase disappears and all CO2 is dissolved
in the oil and brine phases, the aqueous-phase CO2 concentration
increases to 0.0123 m, and the pH drops to 5.62.

CaCO3,solid + H 2 CO3,aq Ca 2+ + 2 HCO3


2
2
aCa 2 + aHCO
aCa 2+ aHCO


K 2HCO 3
3
3
= log10
SI = log10

H 2 CO 3
aCO2 ,aq K sp K1

aCO2 ,aq

K 2HCO 3

+ log10
(T , P ) .
H 2 CO 3
K sp K1

2
2
aCa 2 + aHCO
mCa 2 + mHCO
 Ca 2+ ( HCO )2

3
3
3
=
log

where log10

10
a
m

CO2 ,aq
CO 2, aq
CO2 ,aq

. . . . . . . . . . . . . . . . . . . . . . . . (6)
Scale Prediction
Scale formation tendency can be represented by the thermodynamics-based SI and is defined in the following for calcite:

The SI values for all scale types (e.g., BaSO4 or FeCO3) are calculated by an equation analogous to Eq 5. When SI values at both the
bottomhole and the surface conditions are calculated, SI (= SIsurface
SIbottom hole) can be determined. This SI approach represents the
change in SI vs. (T, P) up the well. If the mineral(s) are at equilibrium in the formation, SIbottomhole = 0. Any calculated deviation from
SIbottomhole = 0.00 is a result of uncertainties in the various measured
parameters, the equilibrium constants, or the calculation procedure
itself. Thereby, most analytical and theoretical uncertainties might
be minimized by using the SI approach. This is often a reliable
way to interpret scale phenomena. Note that the carbonic acid
activity in Eq. 6 can be related to CO2 partial pressure by Henrys

CaCO3,solid Ca 2+ + CO32
aCa 2+ aCO32
Ion Activity Product
SI (Calcite ) Log10
= log10 Calcite
.
K
T
P
)
(
,
K sp (T , P )
sp

. . . . . . . . . . . . . . . . . . . . . . . . (5)
Alkalinity by Titration

Corrected Alkalinity

HDPE

PVS

550

Alkalinity (mg/L)

Alkalinity (mg/L)

PVC

500
450
400
350
300
0
(a)

4
6
Time (Days)

HDPE

550
500
450
400
350
300
0

10
(b)

4
6
Time (Days)

10

Fig. 5Measured Alk of an Fe2+(154 mg/L) containing HCO3 solution (471 mg/L) in 1 M NaCl vs. brine storage time in PVC vs.
HDPE bottles (a) and corrected Alk of the data in Fig. 5a by the procedure (b).
June 2012 SPE Journal

369

SUPCRT92

Eq. 8
9
0

50

100

150

200

250

Barite Log (Ksp)

9.5
10

Activity Coefficients
The Pitzer model is the most accepted ion-interaction model for high
ionic strengths. It has been shown to accurately model the behavior
of electrolyte solution up to 6 mol/kg H2O (Pitzer 1995). At high
TDS, the high ion density of solution can lead to binary interaction
of species of equal and opposite charge and ternary interactions
between three or more ions. The Pitzer model takes into account both
the short-range interactions in concentrated solution and the longrange electrostatic effects of a Dehye-Hkel-type model (Harvie
et al. 1984; Pitzer 1995). The excess free energy is assumed to be
a virial (power series) expansion of binary and ternary interactions
with a leading term of a Debye-Hckel type. Essentially,

10.5
11
11.5
12
Temperature (C)

Fig. 6Comparison of base-10 logarithm of barite solubility


product calculated by Eq. 8 vs. Helgesons Supcrit92 (Johnson
et al. 1992).

law. Therefore, CO2 gas-phase concentration can substitute pH in


calcite SI calculations.
CaCO3,solid + H 2 O + PCO2 Ca 2+ + 2HCO3
2
aCa 2 + aHCO
K 2HCO 3
3
SI Calcite Log10

aCO2,aq
K sp K1HCO 3

aCa 2 + aH2 COK 2HCO 3


3

= Log10
CO2
HCO 3 CO2
P 
K sp K1 K H

CO2 gas

Excess
GPitzer
= RT [ ww f  + ww1 ij ( I )ni n j + ww2 ijk ni n j nk ]
i, j

ijk

with ww = kg of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (9)


where I = ionic strength in molal (m), n is moles, ij and ijk are
binary and ternary constants, and

2
I 1/ 2
f  = A
+
ln(1 + 1.2 I 1/ 2 )
1/ 2
1.2
1 + 1.2 I

with A = 0.3915 at 25 C and 1 atm for water.

. . . . . . . . (7)

where K1HCO 3 and K 2HCO 3 represent the first and second ionization
CO 2
constants of carbonic acid,  gas
represents the fugacity coefficient
CO 2
of CO2 in the gas, and K H represents the Henrys law constant
for the partitioning of CO2 between gas and water. This is the
recommended procedure, when possible, because generally the
partial pressure of CO2 is better known and more easily measured
than the surface brine pH.
Temperature and Pressure Dependence
of Equilibrium Constants
In SI calculations, the temperature and pressure dependence of the
thermodynamic equilibrium constants and the activity coefficients
is needed. Temperature and pressure dependence of equilibrium
constants has been reported extensively by Helgesons group (Helgeson 1967; Tanger and Helgeson 1988; Johnson et al. 1992). The
Helgeson theory is derived from the temperature and pressure
dependence of the partial molar volume and appropriate integration
and differentiation with Maxwells equations to yield the free energy
of formation and the equilibrium constants at any temperature and
pressure. These equations contain 10 to 88 adjustable parameters.
However, at typical oil- and gas-production temperatures and pressures, the Helgeson equations can be greatly simplified because oiland gas-production temperature and pressure are much below the
critical temperature and pressure of water (373.946C and 22 060
kPa, respectively). Most equilibrium constants can be represented
by approximately five adjustable parameters, as has been done by
USGS (Kharaka et al. 1988). For example, Eq. 8 for barite can be
used to replace the Helgeson equations:
Log10Ksp (for Barite) = 136.035 7680.41 / T(K) - 48.595
Log10[T(K)]+[0.394 0.0001119 T(C)] P(atm) / 500.
. . . . . . . . . . . . . . . . . . . . . . . . (8)
370

Fig. 6 is a comparison of logarithmic barite solubility product


[log(Ksp, barite)] calculated by Supcrit92 (Johnson et al. 1992) (on
the basis of the Helgeson theory) vs. that calculated by the simpler
equation (Eq. 8). Excellent agreements are observed between these
two equations in the temperature range of oil and gas production
(0 to 250C).

Then,
G Excess / RT
ln( i ) = Pitzer

ni

T , P ,n j
= Pitzzer activity coefficients after considerablle rearrangement.
. . . . . . . . . . . . . . . . . . . . . . . (10)
For example for a metal (M) and cations ions (c), anions (a),
and neutral molecules (n):
G Excess / RT
ln( M ) = Pitzer
= z m2 F ( I ) + ma (2 BMa + ZC Ma )

nM

a
T , P ,n j
+ mc (2 Mc + ma  Mca )
c

+ ma ma '  Maa '


a < a'

+ z M mc ma Cca + 2 mn nM
c

. . . . . . . . . . . . . . . . . . . . . . . . 11
The Pitzer coefficients BMa, CMa, Mc, Mca, Mca, Cca, and
nM are functions of temperature, pressure, ionic strength, and
composition. For NaCl, a total of 36 constants were included in
Pitzers original derivation, but, for other ions, many parameters
have not been determined. According to Pitzer (1995), the ionpair formation of 2-2 electrolytes (e.g., CaCO30 and CaCO 04 )
has been included in the BMa implicitly.
Pitzer coefficients were derived mostly in binary and ternary systems, such as NaCl-H20 and NaCl-Na2S04-H2O. Several
researchers proposed that an additional explicit ion pair term should
be added if the association constant is greater than 500 for 2-2 type
ion pairs or greater than 20 for 2-l type ion-pairs (Harvie et al. 1984;
He and Morse 1993a, 1993b). Mller (1988) studied the chemical
June 2012 SPE Journal

TABLE 3ESTIMATED ERROR IN CALCITE SI PREDICTION BECAUSE OF ERRORS IN BRINE ANALYSIS

Parameter

Parameter
Initial
Value

Parameter
Final
Value

SI With
Final
Value

(SI / param.)
( SI)/
SI

(Final-Initial)

SI

parameter

Estimated Error
of Parameters
parameter (parameter)

Estimated
2

2 . Error From
SI

param. Each Term


(Final Initial )

T (F)

340

350

0.73139

0.06718

-6.72E-03

4.51E-05

2.50E+01

1.13E-03

P (psia)

7000

7700

0.56316

-0.10105

1.44E-04

2.08E-08

100

1.00E+04

2.08E-04

1.44E-02

19872

20872

0.63995

-0.02426

2.43E-05

5.89E-10

1000

1.00E+06

5.89E-04

2.43E-02

Na (mg/L)
+

K (mg/L)

500

1500

0.64977

-0.01444

1.44E-05

2.09E-10

25

6.25E+02

1.30E-07

3.61E-04

2+

(mg/L)

54

154

0.65621

-0.008

8.00E-05

6.40E-09

2.50E+01

1.60E-07

4.00E-04

2+

(mg/L)

6500

6600

0.67016

0.00595

-5.95E-05

3.54E-09

325

1.06E+05

3.75E-04

1.94E-02

Mg
Ca
Sr

3.36E-02

2+

(mg/L)

700

800

0.66265

-0.00156

1.56E-05

2.43E-10

35

1.23E+03

2.99E-07

5.47E-04

2+

(mg/L)

550

650

0.66308

-0.00113

1.13E-05

1.28E-10

27.5

7.56E+02

9.65E-08

3.11E-04

Fe

2+

(mg/L)

12

20

0.66487

0.00066

-8.25E-05

6.81E-09

4.00E+00

2.72E-08

1.65E-04

Zn

2+

(mg/L)

10

20

0.66403

-0.00018

1.80E-05

3.24E-10

4.00E+00

1.30E-09

3.60E-05

2+

(mg/L)

0.66421

0.00E+00

0.00E+00

4.00E+00

0.00E+00

0.00E+00

Cl (mg/L)

43000

44000

0.67331

0.0091

-9.10E-06

8.28E-11

2150

4.62E+06

3.83E-04

1.96E-02

2-

15

0.66431

1E-04

-1.00E-05

1.00E-10

2.50E+01

2.50E-09

5.00E-05

F (mg/L)

11

0.66407

-0.00014

1.40E-05

1.96E-10

1.00E+00

1.96E-10

1.40E-05

Br (mg/L)

10

20

0.66425

4E-05

-4.00E-06

1.60E-11

2.00E+00

3.20E-11

5.66E-06

SiO2 (mg/L)

10

20

0.66421

0.00E+00

0.00E+00

3.00E+00

0.00E+00

0.00E+00

Alk (mg/L)

281

291

0.7019

0.03769

-3.77E-03

1.42E-05

14.05

1.97E+02

2.80E-03

5.29E-02

10

0.62817

-0.03604

3.60E-03

1.30E-05

10

1.00E+02

1.30E-03

3.60E-02

1.04

1.14

0.63055

-0.03366

3.37E-01

1.13E-01

0.01

1.82E-04

2.06E-05

4.54E-03

1.09E+01

1.19E+02

0.0005

2.50E-07

2.97E-05

5.45E-03

0.66386

-0.00035

3.50E-06

1.22E-11

425

1.81E+05

2.22E-06

50

2.50E+03

8.82E-06

2.97E-03

2.50E+01

7.92E-08

2.82E-04

Ba

Pb
-

SO4 (mg/L)
-

Organic acids
(mg/L)
CO2 (%)

pH=7.1627 pH=7.1240
H2S (%)

0.0283

-0.66421
0.97266

pH=7.1627 pH=7.1722
3

0.30845
-0.66421

Gas (10 SCF)

8500

Oil (BPD)

1000

1100

0.67015

0.00594

-5.94E-05

3.53E-09

100

200

0.66984

0.00563

-5.63E-05

3.17E-09

Brine (BPD)
SI-initial =

8600

0.66421

SUM Sqrs =
Estimated error in Calcite SI value, (SUM Sqrs)

equilibrium model for the quarternary Na-Ca-Cl-SO4,-H20 system


and proposed an explicit ion association constant for CaCO 04 to
be used with Pitzer equations. However, in that authors opinion,
there is not sufficient evidence to support this deviation from the
original Pitzer code. For the most part, the studies by Mller and
others underscore the need for more high-quality experimental data
at high temperature, complex brine and pressure and the need to
closely coupling experimental and modeling efforts, which is the
current undertaking of the authors research group. In the following,
the Pitzer ion interaction model and Helgesons theory are used to
predict the scaling tendency of mineral scales common in oil and
gas production. No standard reference set of Pitzer coefficients has
been established in literature. The authors used a carefully selected
set of Pitzer activity coefficients (Harvie and Weare 1980; Harvie
et al. 1984; Felmy and Weare 1986; Greenberg and Mller 1989; He
and Morse 1993a, 1993b; Kaasa 1998; Monnin 1999; Christov and
Mller 2004; Millero et al. 2007; Duan and Li 2008) to calculate the
scaling index (Eq. 5). Fugacity coefficients for the gas components
are calculated from a Peng-Robinson equation of state. The effect
of oil and gas composition on gas partitioning is estimated from
the Vasquez and Beggs correlation of gas in solution to gas specific
gravity and API oil gravity (Bradley 1987).
Pitzer theory can be modified to predict the effect of hydrate
inhibitors on mineral scaling tendency (Kan et al. 2002; Kan et al.
2003; Kan et al. 2010). For the nonelectrolyte effects [methanol,
MEG or triethylene glycol (TEG)] the SI in the presence of hydrate
inhibitors is defined to be the sum of the values in the absence of an
June 2012 SPE Journal

1/2

6.84E-03
=

0.08270

hydrate inhibitor (i.e., the value calculated by most computer codes)


plus a term that is a function only of the added hydrate inhibitor, SI =
SI (brine only) + SI (because of added MeOH, MEG, or TEG).
In this equation, SI (brine, only) refers to the SI calculated for the
produced brine before any hydrate inhibitor has been added and the
value of SI (because of added MeOH, MEG, or TEG) represents the
effect of added hydrate inhibitor on scale formation at temperature and
ionic-strength. This functional form of temperature and ionic strength
dependence of SI (because of added MeOH, MEG, or TEG) was suggested by Pitzer and others and has been tested by the present authors
and found to represent experimental data on solubility very well (Kan
et al. 2002; Kan et al. 2003; Kan et al. 2010; Lu et al. 2010).
Table 3 lists typical sources of error for calcite SI value using
brine data from a particular gas well in south Texas as an example.
For the purpose of illustration, the brine composition has been
slightly modified to include the minor components. Most cations
can be measured accurately with inductively coupled plasma atomic
emission spectroscopy to within 10% errors. However, some ions
such as Mg2+, K+, Fe2+, and Zn2+ are often not measured in field brine.
Ca/Fe ratio is often approximately 100 if there is no corrosion taking
place (calcite/siderite). Measured pH can easily be off by 0.5 pH
units, and it must be measured at the wellsite. The Alk is often off
by 30 to 100%, unless it is calculated by the method recommended
earlier, which produces both true alkalinity and total carboxylic
acids simultaneously. The overall estimated error in SI can be calculated from the various sources of error, by sum of the squares
and by the square root. Estimated error in each parameter is listed
371

Calcite, 140 F

Barite, 203 F

Halite, 32-536 F

Bypsum, 122 F

Barite

1.0
Anhydrite, 122 F

Celestite, 169 F

0.5
Barite SI

Calculated SI

1.5
1
0.5

0.0
0.5
1 m NaCl

0
1.0

0.5

200

400

Temperature (F)

1.5
2

(b)
0

Ionic Strength

(a)

Barite
372 F

Ellis, 1 m NaCl, 12 atm CO2

478 F

1
0.8
0.6
0.4
0.2
0
0.2
0.4
0.6
0.8
1

1.0
0.5
Calcite SI

Barite SI

205 F

0.0
0.5
1.0

10000

200

20000

Pressure (psia)
(c)

400

600

Temperature (F)
(d)

Fig. 7Plot of Pitzer-theory-based SI of selected literature-reported mineral-salt-solubility data: (a) the calculated SI of selected
solubility data for calcite, barite, halite, gypsum, anhydrite, and celestite at different temperature and ionic strength where the
pressure is typically close to the vapor pressure of water (Ellis 1959; Templeton 1960; Bock 1961; Ellis 1963; Marshall et al.
1964; Blount 1977; Pitzer et al. 1984; Wolf et al. 1989; Howell et al. 1992); (b) barite SI of solubility data at 77 to 400F (Kan et al.
2005); (c) barite SI of solubility data at 200 to 400F and 0- to 15,000-psia pressure (Blount 1977); and (d) calcite SI of solubility
data from 77 to 500F and vapor pressure of water (Ellis 1959, 1963; Wolf et al. 1989).

in Column 9 (parameter). The contribution of each term is listed under


Column 11. On the basis of a simple propagation-of-error formula,
the estimated error in SI = 0.083 [=(SI/Parameter)22parameter]1/2
[i.e., SI (calcite) = 0.664 0.083]. This error estimation neglects systematic errors from equilibrium constants and activity coefficients.
Results have been validated by calculating SI with large numbers
of experimentally measured solubility data (Ellis 1959; Templeton
1960; Bock 1961; Ellis 1963; Marshall et al. 1964; Blount 1977;
Pitzer et al. 1984; Wolf et al. 1989; Howell et al. 1992) at a wide
range of temperature, pressure, and electrolyte compositions (Fig. 7)
(Table 4). Because the solution is assumed to be at equilibrium with
a mineral phase, the theoretical SI for all experiments should be 0.00,
while SI > 0.0 represents a supersaturated solution and SI < 0.0
represents an undersaturated solution. Excellent agreement between
laboratory measured solubilities and the calculated SI values is
observed. Table 4 summarizes the calculated SI for six minerals and
123 solubility data points. The mean SI is 0.013 with a standard
deviation of 0.14, an excellent agreement with the theoretical SI of
0.00, validating the applicability of the scale-prediction model under
these diverse oilfield conditions. However, it should be noted that
the relevant solubility data in complex brines are very limited to
sufficiently test the model in complex, high-TDS brines. Similarly,
there are few high-pressure data to validate the model prediction at
high-pressure/high-temperature (HP/HT) conditions. The authors
research group is currently working on resolving such knowledge
372

gaps. Fig. 8 plots the effects of methanol on salt solubilities and SI


predictions. Hydrate inhibitors, especially methanol, have a strong
impact on mineral-salt solubilities, and the equations presented by
Kan et al. (2002, 2003, 2010) are able to model the effect. The
range of applicability is listed in Table 5. Similarly good agreement
between laboratory observations and predictions is observed.
pH Calculation
For a known solution containing a fixed concentration of sodium
bicarbonate and acetic acid at equilibrium with the CO2 partial
pressure, the solution pH can be calculated from the following
charge-balance equation:
[ Na + ] + [ H + ] [ HCO3 ] 2[CO32 ] [ AC ] [OH ] = 0
[ Na + ] +
2

pH

K
10

H+

CO 2
H

PCO2  g ,CO2 K

H 2 CO 3
1

 HCO

or

10 pH

K HCO 2 PCO2  g ,CO2 K1H 2CO 3 K 2HCO 3

 CO

(10 pH )2

TAc
Kw

= 0.

10 pH  Ac
10 pH  OH
+
1
K 

HAc HAc,aaq

. . . . . . . . . . . . . . (12)

June 2012 SPE Journal

TABLE 4OVERALL QUALITY OF SI PREDICTION


Scale Type

I (m)

T (C)

P (psia)

SI S.D.

Calcite

0 6.24

0 200

14.7 174

0.047 0.121

30

Barite

0 5.0

22 250

14.7 14,417

-0.013 0.147

36

Halite

6.097 10.4

0 300

14.7 1,244

-0.003 0.041

13

Gypsum

0 6.08

0 70

14.7 50

0.036 0.144

18

Anhydrite

0 6.30

150

14.7

-0.087 0.081

Celestite

04

20 150

14.7 7,409

-0.137 0.156

19

Summary Estimated Range of Applicability


All Scales

0 10

0 200

0% MeOH

Gypsum
1.E-02
1.E-03

Barite SI

Ba, Sr, or Ca Conc.


(moles/Kg H2O)

1.E-01

Celestite

1.E-04
1.E-05

Barite
1.E-06
0 .0

0.1

0.2

-0.013 0.14

14.7 15,000

0.3

25% MeOH

2
1.5
1
0.5
0
0.5 0
1
1.5
2

100

0.4

(a)

123

75% MeOH

49% MeOH

200

300

400

Temperature (F)

(b)
Methanol Conc. (Mole Fraction)

25% MeOH

49% MeOH

75% MeOH

Literature
1.0

7
6

25 C

5
Halite SI

NaCl Solubility (mol/Kg H2O)

0% MeOH
Obs.

4
3
2
0

20

40

60

80

0.0
0

100

200

300

400

100

MeOH (wt%)

(c)

1.0
Temperature (F)

(d)

Fig. 8(a) Sulfate salts solubility in methanol/NaCl/water solutions; (b) calculated barite SI using barite solubility data in 075%
methanol/NaCl/water solution at 77 to 400F; (c) literature-reported and observed halite solubility in methanol/NaCl/water solution
at 25C; and (d) calculated halite SI using halite solubility data in 0 to 75% methanol/NaCl/water solution at 77 to 400F.

TABLE 5APPLICABLE RANGE OF SI PREDICTION IN THE PRESENCE OF HYDRATE


INHIBITORS

June 2012 SPE Journal

Mineral

Cosolvent

Cosolvent Conc. (wt%)

I (mol/Kg H2O)

T (C)

Barite

MeOH

0 75

03

0 200

Barite

MEG

0 75

13

0 200

Barite

TEG

0 90

13

0 25

Halite

MeOH

0 90

2.8 6.2

0 200

Halite

MEG

0 80

6.2 8.6

0 200

Celestite

MeOH

0 50

03

0 25

Gypsum

MeOH

0 50

03

25

Calcite

MeOH

0 75

03

0 100

Calcite

MEG

0 75

03

0 100

373

Calc pH (I)

Obs pH (I)

050% MeOH, 00.5 atm PCO2, 00.1 m Ca

Calc pH (II)

Obs pH (II)

075% MeOH, 1 atm PCO2

6.5

2575% MEG, 1 atm PCO2


NaHCO3=16.53 mM, 60C
6.30

I. HAC=8.53 mM

6.0

y = 0.998x

II. HAC=35.18 mM

R = 0.98

6.10
5.5
pH

Calc. pH

5.90

5.0

5.70
5.50

4.5

5.30
4.0

5.3
0

100

200

(a)

300

400

500

5.8

600
(b)

PCO2 (psi)

6.3

Measured pH

Fig. 9pH calculated from Eq. 12 vs. measured pHs. (a) the pH of two solutions at equilibrium with different CO2 partial pressure in the gas phase. The experimental data were from Kaasa (1998). The solutions contained 16.53 mM sodium bicarbonate
and either 8.53 or 35.18 mM acetic acid. The pHs were measured at 60C, and the solution is at equilibrium with 0- to 500-psi
CO2 partial pressure (0 to 40 bar). (b) The pH of calcite-saturated solutions containing various weight% of methanol or MEG at
0 to 1 atm of CO2 partial pressure.

Kaasa (1998) reported the pH values of two solutions containing


16.53 mM NaHCO3 and 8.53 and 35.18 mM acetic acid at equilibrium with 0- to 600-psi CO2 partial pressure at 60C. Fig. 9a
compares the calculated pH values vs. that measured by Kaasa
(1998). The calculated and measured pHs indicate that the equilibrium constants and activity coefcients used in Eq. 12 agree with
measurements (Fig. 9a). Furthermore, the pH values of several
calcite-saturated brine solutions containing various weight% of
either methanol or MEG were measured in the laboratory. These
experiments were performed with a background electrolyte of 1
molal NaCl, 0 to 0.1 M Ca2+, and a CO2 partial pressure of 0- to
1-atm PCO2. Excellent agreement of the measured and predicted
pHs was also observed (Fig. 9b).

Density Calculation
The excess molar volume can be estimated from the pressure derivative
of Pitzer activity coefficients [i.e., Vi ex = log( i P ) RT ]. Therefore,
the density of the salt solution can be estimated from Pitzer theory by

1000 + mi MWi

1000 / H2O + (Vi 0 + Vi ex ) mi

. . . . . . . . . . . . . . . . . . (13)

Excellent agreement has been observed between the calculated


density of 16 concentrated salt solutions and that reported in the
CRC handbook of chemistry and physics (Table 6) as well as the
density of commercial weighting fluid up to 12.7-lbm/gal density
(Fig. 10).

TABLE 6DENSITY CALCULATED BY EQ. 13 VS. THAT REPORTED IN CRC HANDBOOK


OF CHEMISTRY AND PHYSICS (20C)
Density ( D420, g/mL)

Salt Solution

374

Conc. (M)

Reported in CRC
Handbook of
Chemistry and Physics

Density (g/mL)
Calculated by Eq. 13

Deviation
(g/mL)

Acetic acid

6.255

1.043

1.049

0.006

Na Acetate

4.243

1.160

1.170

0.010

KCl

3.742

1.168

1.162

0.006

NaCl

5.326

1.197

1.185

-0.012

MgCl2 6H2O

4.021

1.276

1.287

0.011

BaCl2

1.597

1.279

1.282

0.003

MgSO4 7H2O

2.799

1.296

1.262

-0.034

CaCl2 2H2O

5.03

1.395

1.420

0.026

HCl

6.022

1.098

1.097

-0.001

NaBr

5.495

1.414

1.421

0.007

SrCl2

2.293

1.298

1.303

0.005

H2SO4

5.313

1.303

1.288

-0.015
0.012

KHCO3

2.801

1.169

1.181

K2CO3

4.093

1.414

1.473

0.059

NaHCO3

0.743

1.043

1.042

-0.001

Na2CO3

1.638

1.157

1.170

0.013

Average deviation

0.0060.019

June 2012 SPE Journal

13.0
y = 0.9981x

Calc. Density (lb/gal)

12.5

R2 = 0.998

12.0
11.5
11.0
10.5
10.0
9.5
9.0
8.5
8.0
8.0

8.5

9.0

9.5

10.0

10.5

11.0

11.5

12.0

12.5

13.0

Measured Density (lbm/gal)

Fig. 10Density of commercial weighting fluids calculated by Eq. 13 vs. that measured at 25C.

Minimum Inhibitor Concentration and


Critical SI Prediction
As shown in Fig. 1, scale may not form at low SI. For calcite,
experience has shown that the transition from controllable to
uncontrollable scale formation occurs over a narrow SI range.
The transition SI value from controllable to uncontrollable scale
formation can vary from approximately 2.0 to a maximum of 2.5
for calcite and may depend upon flow conditions, temperature,
specific brine compositions, inhibitor type, and other parameters.
Even though the precise transition SI value is not well defined,
there is almost universal agreement that such a region exists and
care should be taken to avoid any production conditions that will
push SI values into this range of massive and rapid uncontrollable
scale formation. Eqs. 13 through 15 can be used to predict if scaleinhibitor treatment is needed and if so, the minimum inhibitor
concentration (MIC) to inhibit scale. The model is derived from a
large number of nucleation kinetic data of barite, calcium sulfates,

Calcite SI

NTMP
70

60

2.5

50

40
1.5
30

NTMP (mg/L)

Calcite SI

1
20
0.5

10

0
0

50

100

0
150

Temperature (C)
Fig. 11Calculated calcite SI and inhibitor need vs. time. The
brine composition of a gas well in Texas is used for illustration.
June 2012 SPE Journal

and calcite and the inhibition efficiencies of eight scale inhibitors


(He et al. 1996, 1997, 1999; Xiao et al. 2001; Fan et al. 2010).
t0 ( sec ) = 10

(
+
/SI +

binh ( L /mg ) = 10
Cinh ( mg /L ) =

/T ( K ) +
3 /(SI T ( K ))

) . . . . . . . . . . . . . . . . . . (14)

+ 1SI + 2 /T ( K ) + 3 pH + 4 logRinh

) . . . . . . . . . . . (15)

t (sec)
1
log inh
binh (L /mg)
t0 (sec) , . . . . . . . . . . . . (16)

where t0 is the time required for scale to form a precipitate and


tinh is the time required for scale inhibition. On the basis of the
scale-inhibition model, the calcite SI and the MIC at different
temperatures are calculated for a brine composition of a gas well in
Texas and are plotted in Fig. 11. This graph shows an exponential
increase in inhibitor concentration at approximately SI = 2.3 for
calcite, which is in excellent agreement with the field-observed
transition SI for uncontrollable scale formation.
McElhiney et al. (2006) used Eqs. 14 through 16 to calculate
the concentration of Ba2+ that can be mixed with sulfate-reduced
seawater without barite precipitation. Because of the high sulfate
concentration in seawater, barite tolerance is less than 1 mg/L when
an oilfield brine is mixed with seawater. Using sulfate-removal
technology alone, Ba2+ tolerance can be increased by a factor of 40.
Adding inhibitor treatment [20 mg/L diethylenetriaminepentaphosphonic acid (DTPMP)], Ba2+ tolerance can be increased by a factor
of 100 (Fig. 12). Such effects can be predicted and are validated
experimentally (Fig. 12 solid dots are experimental data).
Conclusions
Scale prediction and control are complex in nature. This paper
summarizes the state of the art in scale prediction and control for
oil and gas application. The following conclusions can be made:
1. Accurate pH is difficult to measure because of both theoretical and practical limitations. If gas-phase CO2 concentration is
known, it should always be used in place of measured pH.
2. Alk measurements carry a large margin of error, unless the
simultaneous Alk and organic-acids titration method discussed
in this paper is used.
3. Scale predictions based on the Pitzer ion-interaction model agree
well with literature solubility data for a temperature range of 0
to 200C, a pressure range of 0 to 15,000 psi, and a TDS range
of 0 to 350,000 mg/L with a standard error of 0.1 SI unit.
4. There is not enough literature data at temperatures greater than
200C, at pressures greater than 15,000 psia, and in complex
375

No Inh

1 mg/L DTPMP

5 mg/L DTPMP

10 mg/L DTPMP

20 mg/L DTPMP

Exp Obs

Ba in Formation Water (mg/L)

10000

1000

100

10

0.1
0

500

1000

1500

2000

2500

3000

SO4 in Sulfate Reduced Seawater (mg/L)

Fig. 12Plot of barium vs. sulfate concentrations that can be tolerated for 120 minutes without precipitation in the presence 0
to 20 mg/L of a scale inhibitor [diethylenetriaminepentaphosphonic acid, (DTPMP)]. The results calculated with Eqs. 14 through
16 were compared to the laboratory-observed results where 10 or 20 mg/L DTPMP was used, and the concentrations are labeled
in parentheses in the plot.

brine of high TDS to validate scale predictions, and more


research is needed in these HP/HT ranges.
5. A modification of the Pitzer theory can be used to predict the
mineral scaling tendency in the presence of methanol and MEG.
Similarly, the same algorithm can be used to calculate the brine
pH, in the presence and absence of hydrate inhibitors, with
excellent precision and to calculate the density of fluids up to
12.7 lbm/gal.
Acknowledgments
This work was financially supported by a consortium of companies
including Baker-Petrolite, BP, Champion Technologies, Chevron,
ConocoPhillips, Dow Chemicals, Halliburton, Hess, Innovacion e
Ingenieria Sustentable S.A. de C.V. Kemira, Marathon Oil, M-I
Swaco, Multi-Chem, Nalco, Occidental Oil and Gas, Petrobras,
Saudia Aramco, Shell, Statoil, Siemens, and Total, and the National
Science Foundation through the Center for Biological and Environmental Nanotechnology [EEC-0118007], the China-US Center
for Environmental Remediation and Sustainable Development, and
the Advanced Energy Consortium.
Nomenclature
ai = ion activity
Alk = alkalinity, molal
Alk0 = initial alkalinity in brine, molal
A = Debye-Hckel constant for osmotic pressure
binh = inhibition-efciency coefcient
BMa, CMa, Mc, Mca, Mca, Cca, and nM = Pitzer second and third
virial coefcients
Cinh = minimum inhibitor concentration, mg/L
Excess
GPitzer
= excess freee energy
I = the ionic strength, molal
Ksp = solubility product
K HCO 2 = Henrys law constant for CO2 partition in gas/water
phase, molal atm1
K1H 2 CO 3 = the rst ionization constant of carbonic acid
K 2HCO 3 = the second ionization constant of carbonic acid
K HAc = the dissociation constant of acetic acid
mo = mass of oil, kg
mw = Ww = mass of water, kg
CO 2
= CO2 in all phases, moles
ntotal
ngCO 2 = CO2 in gas phase, moles
376

nwCO 2

nwHCO3
2
nwCO3
Ni
PCO2
R
Rinh
SI
t0
tinh
T
TAc
Vg
Z

I
I

 CO2, g
CO,o
i
pHj
ij
ijk

=
=
=
=
=
=
=
=
=
=
=
=

CO2 in aqueous phase, moles


bicarbonate in aqueous phase, moles
carbonate in aqueous phase, moles
moles of components
partial pressure of CO2, atm
the gas constant
lattice cation to anion molar ratio
saturation index
nucleation time, seconds
inhibition time, seconds
temperature, K
total concentration of all acetate species, molal =
HAc+Ac
= gas volume, L
= gas compressibility
= parameters to nucleation time
= parameters to predict inhibition efciency
= CO2 fugacity coefcient in gas
= CO2 activity coefcient in oil
= Pitzer activity coefcient
= the change in junction potential in unit of pH
= Pitzer binary constants
= Pitzer ternary constant

References
Alsaiari, H.A., Kan, A., and Tomson, M.B. 2010. Effect of Calcium and
Iron (II) Ions on the Precipitation of Calcium Carbonate and Ferrous
Carbonate. SPE J. 15 (2): 294300. SPE-121553-PA. http://dx.doi.
org/10.2118/121553-PA.
Bard, A.J. and Faulkner, L.R. 1980. Electrochemical Methods: Fundamentals and Applications. New York: John Wiley & Sons.
Bates, R.C. 1973. Determination of pH-Theory and Practice, second edition. New York: John Wiley & Sons.
Blount, C.W. 1977. Barite solubilities and thermodynamic quantities up to
300C and 1400 bars. Am. Mineral. 62: 942957.
Bock, E. 1961. On the Solubility of Anhydrous Calcium Sulphate and of
Gypsum in Concentrated Solutions of Sodium Chloride at 25 C, 30
C, 40 C, and 50 C. Can. J. Chem. 39 (9): 17461751. http://dx.doi.
org/doi:10.1139/v61-228.
Breng, R., Schmidt, T., Vikane, O. et al. 2003. Downhole Measurement
of pH in Oil & Gas Applications by Use of a Wireline Tool. Paper SPE
June 2012 SPE Journal

82199 presented at the SPE European Formation Damage Conference,


The Hague, 1314 May. http://dx.doi.org/10.2118/82199-MS.
Bradley, H.B. 1987. Petroleum Engineering Handbook. Richardson, Texas:
SPE.
Burke, P.A. and Hausler, R.H. 1985. Assessment of CO2 Corrosion in
the Cotton Valley Limestone Trend. In Advances in CO2 Corrosion:
Selected Papers from the Corrosion/84 Symposium on Corrosion by
Co2 in the Oil and Gas Industry and Corrosion/85 Symposium, ed. P.A.
Burke, A.I. Asphahani, and B.S. Wright. Houston, Texas: NACE.
Carothers, W.W. and Kharaka, Y.K. 1978. Aliphatic acid anions in oil-field
waters; implications for origin of natural gas. AAPG Bull. 62 (12):
24412453.
Christov, C. and Mller, N. 2004. Chemical equilibrium model of solution
behavior and solubility in the H-Na-K-OH-Cl-HSO4-SO4-H2O system
to high concentration and temperature. Geochim. Cosmochim. Acta 68
(6): 13091331. http://dx.doi.org/10.1016/j.gca.2003.08.017.
Collins, A.G. and Davis, J.W. 1971. Solubility of barium and strontium
sulfates in strong electrolyte solutions. Environ. Sci. Technol. 5 (10):
10391043. http://dx.doi.org/10.1021/es60057a007.
Deer, W.A., Howie, R.A., and Zussman, J. 1983. An Introduction to the Rock
Forming Minerals. Essex, UK: Halsted Press/John Wiley & Sons.
Duan, Z. and Li, D. 2008. Coupled phase and aqueous species equilibrium of the H2O-CO2-NaCl-CaCO3 system from 0 to 250 C, 1 to
1000 bar with NaCl concentrations up to saturation of halite. Geochim. Cosmochim. Acta 72 (20): 51285145. http://dx.doi.org/10.1016/
j.gca.2008.07.025.
Dunlop, A.K., Hassell, H.L., and Rhodes, P.R. 1985. Fundamental considerations in sweet gas well corrosion. In Advances in CO2 Corrosion:
Selected Papers from the Corrosion/84 Symposium on Corrosion by
Co2 in the Oil and Gas Industry and Corrosion/85 Symposium, ed. P.A.
Burke, A.I. Asphahani, and B.S. Wright. Houston, Texas: NACE.
Ellis, A.J. 1959. The solubility of calcite in carbon dioxide solutions. Am.
J. Sci. 257 (5): 354365. http://dx.doi.org/10.2475/ajs.257.5.354.
Ellis, A.J. 1963. The solubility of calcite in sodium chloride solutions
at high temperatures. Am. J. Sci. 261 (3): 259267. http://dx.doi.
org/10.2475/ajs.261.3.259.
Fan, C., Kan, A., Fu, G., Tomson, M., and Shen, D. 2010. Quantitative
Evaluation of Calcium Sulfate Precipitation Kinetics in the Presence
and Absence of Scale Inhibitors. SPE J. 15 (4): 977988. SPE-121563PA. http://dx.doi.org/10.2118/121563-PA.
Felmy, A.R. and Weare, J.H. 1986. The prediction of borate mineral
equilibria in natural waters: Application to Searles Lake, California. Geochim. Cosmochim. Acta 50 (12): 27712783. http://dx.doi.
org/10.1016/0016-7037(86)90226-7.
Felmy, A.R., Rai, D., and Amonette, J.E. 1990. The solubility of barite and
celestite in sodium sulfate: Evaluation of thermodynamic data. J. Solution Chem. 19 (2): 175185. http://dx.doi.org/10.1007/bf00646611.
Frenier, W.W. and Ziauddin, M. 2008. Formation, Removal and Inhibition
of Inorganic Scale in the Oilfield Environment. Richardson, Texas:
SPE.
Greenberg, J. and Tomson, M. 1992. Precipitation and dissolution kinetics and
equilibria of aqueous ferrous carbonate vs temperature. Appl. Geochem.
7 (2): 185190. http://dx.doi.org/10.1016/0883-2927(92)90036-3.
Greenberg, J.P. and Mller, N. 1989. The prediction of mineral solubilities
in natural waters: A chemical equilibrium model for the Na-K-Ca-ClSO4-H2O system to high concentration from 0 to 250C. Geochim.
Cosmochim. Acta 53 (10): 25032518. http://dx.doi.org/10.1016/00167037(89)90124-5.
Harvie, C.E. and Weare, J.H. 1980. The prediction of mineral solubilities in
natural waters: the NaKMgCaClSO4H2O system from
zero to high concentration at 25 C. Geochim. Cosmochim. Acta 44 (7):
981997. http://dx.doi.org/10.1016/0016-7037(80)90287-2.
Harvie, C.E., Mller, N., and Weare, J.H. 1984. The prediction of mineral solubilities in natural waters: The Na-K-Mg-Ca-H-Cl-SO4-OHHCO3-CO3-H2O-system to high ionic strengths at 25C. Geochim.
Cosmochim. Acta 48 (4): 723751. http://dx.doi.org/10.1016/00167037(84)90098-X.
He, S. and Morse, J.W. 1993a. Prediction of halite, gypsum, and anhydrite
solubility in natural brines under subsurface conditions. Comput. Geosci.
19 (1): 122. http://dx.doi.org/10.1016/0098-3004(93)90039-8.
He, S. and Morse, J.W. 1993b. The carbonic acid system and calcite solubility in aqueous Na-K-Ca-Mg-Cl-SO4 solutions from 0 to
June 2012 SPE Journal

90C. Geochim. Cosmochim. Acta 57 (15): 35333554. http://dx.doi.


org/10.1016/0016-7037(93)90137-l.
He, S.L., Kan, A.T., and Tomson, M.B. 1996. Mathematical Inhibitor
Model for Barium Sulfate Scale Control. Langmuir 12 (7): 19011905.
http://dx.doi.org/10.1021/la950876x.
He, S.L., Kan, A.T., and Tomson, M.B. 1999. Inhibition of calcium carbonate precipitation in NaCl brines from 25 to 90C. Appl. Geochem. 14
(1): 1725. http://dx.doi.org/10.1016/S0883-2927(98)00033-X.
He, S.L., Kan, A.T., Tomson, M.B., and Oddo, J.E. 1997. A New Interactive
Software for Scale Prediction, Control, and Management. Paper presented
at the SPE Annual Technical Conference and Exhibition, San Antonio,
Texas, USA, 58 October. http://dx.doi.org/10.2118/38801-MS.
Helgeson, H.C. 1967. Thermodynamics of complex dissociation in aqueous
solution at elevated temperatures. The Journal of Physical Chemistry 71
(10): 31213136. http://dx.doi.org/10.1021/j100869a002.
Howell, R.D., Raju, K., and Atkinson, G. 1992. Thermodynamics of
scale mineral solubilities. 4. Experimental measurements of strontium sulfate(s) in water and aqueous sodium chloride from 25 to
250C and from 1 to 500 bar. J. Chem. Eng. Data 37 (4): 464469.
http://dx.doi.org/10.1021/je00008a020.
Ikeda, A., Ueda, M., and Mukai, S. 1985. Co2 behavior of carbon and Cr
steels. In Advances in CO2 Corrosion: Selected Papers from the Corrosion/84 Symposium on Corrosion by Co2 in the Oil and Gas Industry
and Corrosion/85 Symposium, ed. P.A. Burke, A.I. Asphahani, and B.S.
Wright. Houston, Texas: NACE.
Johnson, J.W., Oelkers, E.H., and Helgeson, H.C. 1992. SUPCRT92: A
software package for calculating the standard molal thermodynamic
properties of minerals, gases, aqueous species, and reactions from 1 to
5000 bar and 0 to 1000C. Comput. Geosci. 18 (7): 899947. http://
dx.doi.org/10.1016/0098-3004(92)90029-q.
Kaasa, B. 1998. Prediciton of pH, mineral precipitations and multiphase
equilibria during oil recovery. PhD Thesis No. IUK 103, Department
of Materials Science and Engineering, NTNU, Trondheim, Norway,
267.
Kan, A.T., Fu, G., and Tomson, M.B. 2002. Effect of Methanol on Carbonate Equilibrium and Calcite Solubility in a Gas/Methanol/Water/
Salt Mixed System. Langmuir 18 (25): 97139725. http://dx.doi.
org/10.1021/la025620n.
Kan, A.T., Fu, G., and Tomson, M.B. 2003. Effect of Methanol and Ethylene Glycol on Sulfates and Halite Scale Formation. Ind. Eng. Chem.
Res. 42 (11): 23992408. http://dx.doi.org/10.1021/ie020724e.
Kan, A.T., Fu, G., and Tomson, M.B. 2005. Prediction of Scale Inhibitor Squeeze and Return in Calcite-Bearing Formation. Paper SPE
93265 presented at the SPE International Symposium on Oilfield
Chemistry, The Woodlands, Texas, USA, 24 February. http://dx.doi.
org/10.2118/93265-MS.
Kan, A.T., Lu, H., and Tomson, M.B. 2010. Effects of Monoethylene
Glycol on Carbon Dioxide Partitioning in Gas/Monoethylene Glycol/
Water/Salt Mixed Systems. Ind. Eng. Chem. Res. 49 (12): 58845890.
http://dx.doi.org/10.1021/ie901274v.
Kharaka, Y.K., Gunter, W.D., Aggarwal, P.K., Perkins, E., and DeBraal,
J.D. 1988. SOLMINEQ.88A computer program for geochemical
modeling of water-rock reactions. Water-Resources Investigations
Report 88-4227, USGS, Menlo Park, California.
Lu, H., Kan, A.T., and Tomson, M.B. 2010. Effects of Monoethylene
Glycol on Carbonate Equilibrium and Calcite Solubility in Gas/Monoethylene Glycol/NaCl/Water Mixed Systems. SPE J. 15 (3): 714725.
SPE-121562-PA. http://dx.doi.org/10.2118/121562-PA.
Lundegard, P.D. and Kharaka, Y.K. 1990. Geochemistry of organic acids
in subsurface waters. Field data, experimental data, and models. In
Chemical Modeling of Aqueous Systems II, ed. D.C. Melchior and R.L.
Bassett, No. 416, 169189. Washington, DC: ACS Symposium Series,
American Chemical Society.
Lundegard, P.D. and Land, L.S. 1993. Carboxylic acid anions in formation
waters, San Joaquin Basin and Louisiana Gulf Coast, U.S.A.Implications for clastic diagenesis. Discussion. Appl. Geochem. 8 (3): 297300.
http://dx.doi.org/10.1016/0883-2927(93)90044-h.
Marshall, W.L., Slusher, R., and Jones, E.V. 1964. Aqueous Systems at
High Temperatures XIV. Solubility and Thermodynamic Relationships
for CaSO4 in NaCl-H2O Solutions from 40 to 200C, 0 to 4 Molal
NaCl. J. Chem. Eng. Data 9 (2): 187191. http://dx.doi.org/10.1021/
je60021a011.
377

McElhiney, J., Kan, A.T., and Tomson, M.B. 2006. Design of Low Sulfate
Seawater Injection Based Upon Kinetic Limits. Paper presented at the
SPE International Oilfield Scale Symposium, Aberdeen, 31 May1
June. http://dx.doi.org/10.2118/100480-MS.
Millero, F., Huang, F., Graham, T., and Pierrot, D. 2007. The dissociation
of carbonic acid in NaCl solutions as a function of concentration and
temperature. Geochim. Cosmochim. Acta 71 (1): 4655. http://dx.doi.
org/10.1016/j.gca.2006.08.041.
Mller, N. 1988. The prediction of mineral solubilities in natural waters:
A chemical equilibrium model for the Na-Ca-Cl-SO4-H2O system, to
high temperature and concentration. Geochim. Cosmochim. Acta 52 (4):
821837. http://dx.doi.org/10.1016/0016-7037(88)90354-7.
Monnin, C. 1999. A thermodynamic model for the solubility of barite and celestite in electrolyte solutions and seawater to 200C and to 1 kbar. Chem. Geol.
153 (14): 187209. http://dx.doi.org/10.1016/s0009-2541(98)00171-5.
Monnin, C. and Galinier, C. 1988. The solubility of celestite and barite
in electrolyte solutions and natural waters at 25C: A thermodynamic
study. Chem. Geol. 71 (4): 283296. http://dx.doi.org/10.1016/00092541(88)90055-1.
Okocha, C. and Sorbie, K.S. 2010. Effects of Sulphide Scales (PbS, ZnS
& FeS) on BaSO4 Crystal Growth and Dissolution. Paper SPE 130391
presented at the SPE International Conference on Oilfield Scale, Aberdeen, 2627 May. http://dx.doi.org/10.2118/130391-MS.
Okocha, C., Sorbie, K.S., and Boak, L.S. 2008. Novel Inhibition Mechanism
for Sulfide Scales. Paper SPE 112538 presented at the SPE International
Symposium and Exhibition on Formation Damage Control, Lafayette,
Louisiana, USA, 1315 February. http://dx.doi.org/10.2118/112538-MS.
Pitzer, K.S. 1995. Thermodynamics, third edition. New York: McGraw Hill
Series in Advanced Chemistry, McGraw-Hill.
Pitzer, K.S., Peiper, J.C., and Busey, R.H. 1984. Thermodynamic Properties
of Aqueous Sodium Chloride Solutions. J. Phys. Chem. Ref. Data 13
(1): 1102. http://dx.doi.org/10.1063/1.555709.
Raghuraman, B., OKeefe, M., Eriksen, K.O., et al. 2007. Real-Time Downhole pH Measurement Using Optical Spectroscopy. SPE Res Eval & Eng
10 (3): 302311. SPE-93057-PA. http://dx.doi.org/10.2118/93057-PA.
Reardon, E.J. and Armstrong, D.K. 1987. Celestite (SrSO4(s)) solubility in
water, seawater and NaCl solution. Geochim. Cosmochim. Acta 51 (1):
6372. http://dx.doi.org/10.1016/0016-7037(87)90007-x.
Shen, D., Fu, G., Al-Saiari, H.A., Kan, A.T., and Tomson, M.B. 2009.
Barite Dissolution/Precipitation Kinetics in Porous Media and in the
Presence and Absence of a Common Scale Inhibitor. SPE J. 14 (3):
462471. SPE-114062-PA. http://dx.doi.org/10.2118/114062-PA.

378

Sontheimer, H., Kolle, W., and Snoeyink, V. 1981. The Siderite Model of the
Formation of Corrosion-Resistant Scales. JAWWA 71 (11): 572559.
Sun, W., Neic, S., and Papavinasam, S. 2008. Kinetics of Corrosion Layer
Formation. Part 2Iron Sulfide and Mixed Iron SulfideCarbonate Layers in Carbon DioxideHydrogen Sulfide Corrosion. Corros. Sci. 64 (7):
586599. http://dx.doi.org/10.5006/1.3278494.
Tanger, J.C. and Helgeson, H.C. 1988. Calculation of the thermodynamic
and transport properties of aqueous species at high pressures and
temperatures; revised equations of state for the standard partial molal
properties of ions and electrolytes. Am. J. Sci. 288 (1): 1998. http://
dx.doi.org/10.2475/ajs.288.1.19.
Templeton, C.C. 1960. Solubility of Barium Sulfate in Sodium Chloride
Solutions from 25 to 95C. J. Chem. Eng. Data 5 (4): 514516. http://
dx.doi.org/10.1021/je60008a028.
Tomson, M.B. and Johnson, M.L. 1991. How Ferrous Carbonate Kinetics
Impacts Oilfield Corrosion. Paper SPE 21025 presented at the SPE
International Symposium on Oilfield Chemistry, Anaheim, California,
USA, 2022 February. http://dx.doi.org/10.2118/21025-MS.
Tomson, M.B. and Oddo, J.E. 1996. Handbook of Calcite Scale Measurement, Prediction and Control. Houston, Texas: Hart Publications, Inc.
Tomson, M.B., Kan, A.T., Fu, G., and Cong, L. 2006. Measurement of
total Alkalinity and Carboxylic Acid and Their Relation to Scaling
and Corrosion. SPE J. 11 (1): 103110. SPE-87449-PA. http://dx.doi.
org/10.2118/87449-PA.
Wolf, M., Breitkopf, O., and Puk, R. 1989. Solubility of calcite in different
electrolytes at temperatures between 10 and 60C and at CO2 partial
pressures of about 1 kPa. Chem. Geol. 76 (3-4): 291301. http://dx.doi.
org/10.1016/0009-2541(89)90097-1.
Xiao, J., Kan, A.T., and Tomson, M.B. 2001. Prediction of BaSO4 Precipitation in the Presence and Absence of a Polymeric Inhibitor: Phosphinopolycarboxylic Acid. Langmuir 17 (15): 46684673. http://dx.doi.
org/10.1021/la001721e.
Amy Kan is the codirector of the Brine Chemistry Consortium at
Rice University. She holds a PhD degree from Cornell University.
She serves as the associate editor for SPEJ, and as member
of the executive committee on SPE International Oilfield
Scale Conference and SPE International Nanotechnology
Conference.
Mason B. Tomson has been a professor of civil and environmental engineering at Rice University since 1987. He holds a PhD
degree in physical chemistry from Oklahoma State University.

June 2012 SPE Journal

Vous aimerez peut-être aussi