You are on page 1of 11

Hot Oil and Gas Wells Can Be Stimulated

Without Acids
Wayne Frenier, SPE, Mark Brady, SPE, Salah Al-Harthy, SPE, Roberto Arangath, Keng Seng Chan, SPE,
Nicolas Flamant, SPE, and Mathew Samuel, SPE, Schlumberger

Summary
A revolutionary family of treating fluids designed for the stimulation of critical, hot, or exotic oil and gas wells has been developed through application of detailed chemical and engineering
studies.13 Formulations based on the hydroxethylaminocarboxylic acid (HACA) family of chelating agents have now been used
to successfully increase production of oil and gas from wells in
a variety of different formations. Included in the field test matrixes were new and producing wells drilled into carbonates
and sandstone formations. The temperatures of the wells treated
ranged from 230 to 370F (110 to 187C) bottomhole static temperature (BHST).
Because these formulations do not contain high concentrations
of corrosive mineral or organic acids (the formulations are less
acidic than carbonated beverages), very low corrosion rates of the
tubulars can be achieved by application of small amounts of special, inexpensive corrosion inhibitors. The mild fluids also are
highly retarded so that high-temperature carbonates can be stimulated and sensitive sandstone formations are not damaged. The
fluids have reduced health, safety, and environmental (HSE) footprints because:
(1) They are much less toxic to mammals as well as to aquatic
organisms than mineral acids or organic acids such as hydrochloric
(HCl), hydrofluoric (HF), or formic acid.
(2) The fluids are returned to the surface at pH values between
5 and 7, and they frequently can be added to normal well production fluids without neutralization.
(3) Because of much lower corrosion rates for corrosion resistant alloys (CRAs), lowered concentrations of Ni and Cr will be in
the well returns compared with conventional acids that also may
contain antimony (as a corrosion inhibitor).
Introduction
While mineral acids can be very effective stimulation fluids at low
temperatures, the use of HCl-based fluids at high temperatures
[generally defined as greater than 200F (93C)] can cause many
problems. The major concerns are damage to corrosion-resistant
tubular materials, toxicity of the fluids and inhibitors, too rapid attack
on the formation (carbonates), and massive damage to clays in sandstone formations. Alternative fluids based on the HACA family of
chelating agents can be formulated to alleviate these problems.
This paper will describe the scientific basis for using these
fluids in hot formations. We also describe a completely new family
of matrix stimulation fluids, based on HACA chemicals, that has a
unique ability to be tailored to specific formation conditions. Because of the high acid solubility of HACA chemicals, formulations
of low- as well as high-pH fluids have been produced. A major
application will be that of stimulating high-temperature carbonate
formations where mineral acids cannot be pumped fast enough to
produce wormholes unless these are retarded by the formation of
emulsions. In addition, this paper describes results from laboratory
tests and field treatments using chelating agent fluids for matrix

Copyright 2004 Society of Petroleum Engineers


This paper (SPE 86522) was first presented at the 2004 SPE International Symposium and
Exhibition on Formation Damage Control, Lafayette, Louisiana, 1820 February, and revised for publication. Original manuscript received for review 18 June 2003. Revised manuscript received 5 August 2003. Paper peer approved 17 August 2004.

November 2004 SPE Production & Facilities

stimulation of high-temperature sandstone formations. Laboratory


experiments have been conducted up to 400F (204C) and have
included rotating disk tests using carbonate specimens to determine the kinetics and coreflood tests using carbonate and sandstone cores to validate dissolution mechanisms and to qualify formulations for use in field applications. Results from field applications up to 370F (187C) are presented.
Literature on Use of Chelating Agents in Well Stimulation.
Chelating agents are materials used to control undesirable reactions of metal ions. In oilfield chemical treatments, chelating
agents1 are frequently added to stimulation acids to prevent precipitation of solids as the acid spends on the formation being
treated. See references by Frenier2 and Frenier et al.3 for more
detailed reviews. The materials, which were evaluated, include
HACA such as hydroxyethylethylenediaminetriacetic acid
(HEDTA) and hydroxyethyliminodiacetic acid (HEIDA), as well
as other types of chelating agents.
Fredd and Fogler46 have proposed uses for ethylenediaminetetraacetic acid (EDTA)-type chelating agents. This application
uses the chelating agents as the primary dissolution agent in matrix
acidizing of carbonate formations [calcite, which is calcium
(CaCO3) carbonate, and dolomite, which is calcium/magnesium
carbonate(Ca/MgCO3)]. Because HCl reacts so rapidly on most
carbonate surfaces, diverting agents, ball sealers, and foams7 are
used to direct some of the acid flow away from large channels that
may form initially and take all the subsequent acid volume. By
adjusting the flow rate and pH of the fluid, it may be possible to
tailor the slower-reacting chelate solutions to the well conditions
and achieve maximum wormhole formation with a minimum
amount of solvent.
Disodium EDTA has been used as a scale-removal agent in
the Prudhoe Bay field of Alaska.8,9 In these applications, CaCO3
scale had precipitated in the perforation tunnels and in the nearwellbore region of a sandstone formation. Huang et al.10 described organic acid formulations for removal of scale and fines at
high temperatures.
One aspect of chelating agent fluids has proven to be most
useful for treating a wide range of formations and damage mechanisms. This is the large range of different types of formulations
that can be produced by changing the pH with addition of acids or
bases. The most common commercial fluids available are tetrasodium EDTA and trisodium HEDTA; these have pH values of
approximately 12. Table 1 shows the pKa values for the carboxylate groups in these molecules. These values also define the
buffer points because the buffer power is at a maximum when
pHpKa. Many different formulations (usually proprietary) can
be produced by addition of mineral acids or organic acids to sodium EDTA or sodium HEDTA to make acidic fluids that are quite
aggressive for dissolving calcite. Based on the pK values, HEDTA
would buffer strongly at pH 2.6 and 5.4 (measured at 25C), while
EDTA could buffer at pH 2.0, 2.7, and 6.1. However, only
HEDTA fluids can actually be produced as formulation with pH
values <5.0 because of the much higher solubility of HEDTA acid
compared with EDTA acid.
Experimental Procedures
The experimental program included tests to determine the kinetic
parameters for dissolution of calcite using the rotating disk methods and for determining the extent of wormhole formation using
coreflood tests.
189

The procedures for performing the rotating disk tests (using


marble) and limestone coreflood tests, including the composition
of the disks and cores, are described by Frenier et al.3,11 The cores
were photographed and reweighed after the test was completed.
The cores were imaged using computed tomographic (CT) X-ray
scans.12 The scans were run at 130 kV on dry cores. The structures
of the chelating agents are shown in Frenier et al.,11 while the
stability constants for the materials are in Table 1 of this paper.
These materials were used as aqueous solutions.
Coreflood tests3 were also run using Berea sandstone cores.
The Berea cores, 16 in. (25150 mm), were treated with several
different solvents. X-ray diffraction (XRD)-determined compositions are in Table 2. The effluents were collected and analyzed using inductively coupled plasma optical emission spectroscopy (ICPOES).
Dissolution reactions for various clay minerals were carried out
using a slurry reactor described by Hartman et al.13 The Hastelloy
B reactor (1L) was lined with polytetrafluoroethylene and was
stirred at a constant rate of 100 rpm during each experiment. The
grain size of the minerals used in the study was approximately 100
m. Samples (2 to 4 mL) were withdrawn from the reactor at
various time intervals and filtered using syringes fitted with membrane filters with a pore size of 0.45 mm. Immediately after filtration, dilute nitric acid was added to the filtrate to prevent precipitation. The filtrate samples were then analyzed for dissolved
aluminum, silicon, and other metals with ICPOES in order to
monitor the reaction progress.
Corrosion weight-loss tests were conducted using hightemperature autoclaves.2 One inhibitor, (A), was formulated to
be compatible with mineral acids, and another inhibitor, (B),
was formulated to be compatible with organic acids and chelating agents.

Results and Discussion


Calcite Tests. The work described by Frenier et al.11 is summarized herein. Rotating disk tests with marble were performed with
HEDTA fluids. Eq. 1 describes the flux (J) to the rotating disk as
a function of the rotation rate (), diffusion coefficient (Deff), and
the difference in the concentration gradient between the interface
(I) and the bulk (b). The surface reaction rate coefficient [kr
kyKy/(1+KyCI)] is a function of adsorption equilibrium and a reaction rate constant.4 It was assumed to be constant, and a pseudofirst-order rate expression (Eq. 2) was used to apply the definition
of the Damkhler number.
J = Rate = 0.62 Deff2 3 1 2Cb C1, . . . . . . . . . . . . . . (1)
Rate = krCI. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
Diffusion coefficients were estimated based on 0.001-m data and
the Arrhenius equation with Ea4 kcal/mol. The diffusion coefficients calculated using these data and Eq. 1 are also in agreement
with the literature14,15 for dilute solutions (0.001m) on marble.
Additional reaction rate parameters were developed using 0.65-m
solutions of trisodium HEDTA (Na3HEDTA).
Fig. 1 shows the surface reaction rates for marble (600 rpm) as
a function of temperature in 0.65-m HEDTA at three pH values.
The calculated activation energy (Ea) values were 12.4 to 13.2
kcal/mol. The values are in the surface reaction rate controlled
range. The activation energy values are consistent with data reported for sodium EDTA formulations (pH 4 to 12), where there
are comparable data.5
Work by Fredd and Fogler4,6 demonstrated that the effects of
mass transport and, to a lesser extent, the surface reaction rate
dominate wormhole formation during stimulation of carbonate
rocks. In their work, the Damkhler number (NDa, the ratio of the
overall reaction rate/convective transport) is used to define the
various regions of wormhole formation. The Damkhler number is
given by
NDa = dl q. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)
1+

1
vKeq

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (4)
1
1
1
+
+
K1 vkr vKeqK3

Kmt = 1.86Deff2 34q d3l1 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . (5)


Here, Deffthe effective diffusion coefficient for the active species, lthe length of the wormhole, qthe flow rate, krthe
surface reaction rate constant, and dthe diameter of the wormhole. The term the overall dissolution rate constant. At high
190

November 2004 SPE Production & Facilities

Fig. 1Temperature dependency of marble dissolution in 0.65 m Na3HEDTA.

Damkhler numbers (low injection rates), the reactants are consumed at the core face, and little penetration (wormhole formation)
is achieved. The Damkhler number for a cylindrical wormhole
can be calculated using Eqs. 3 through 5. The Fredd and Fogler46
papers propose that the optimal Damkhler number is approximately 0.29 for a large number of solvents. Using this value and an
average wormhole diameter of 0.16 cm, the optimal flow rates for
Na3HEDTA fluids with pH of 12, 4, and 2.5 were calculated and
displayed as a function of temperature (Fig. 2).
To validate these predictions, coreflood tests with Indiana limestone were performed with various chelating agents and acid systems from 150 to 400F [65 to 204C]. All the HEDTA-based
fluids stimulated the cores; some produced as much as a 400-fold
increase in permeability.
The CT scans and inlet photographs of the wormholes for tests
run at 250F (121C) and 2 mL/min are shown in Fig. 3. Wormholes were observed to penetrate the entire length of the cores for
all pH values. The diameter of the wormhole formed with the
aggressive HEDTA pH 3.5 with formic acid fluid was much larger
than that formed with the other HEDTA fluids. This observation is
consistent with the weight loss of the core.
Several HEDTA formulations were tested at 350F (177C). In
addition, 10% acetic acid was used to stimulate the limestone core.
This acid was tested because it is currently one of the few materials
used at these high temperatures. A new procedure was used to
relate the laboratory core flow data to radial flow in a limestone
formation. The procedure relied on previous work performed by
Fredd and Fogler,16 who normalized the wormhole formation phenomenon with a single 3D surface. The kinetic and equilibrium
constants measured for HEDTA were incorporated into this model.
In addition, computational fluid dynamics (CFD) simulations were
performed with a commercial simulator to investigate the effects
of the pressure and velocity fields on the wormhole density in
radial flow. Fig. 4 shows how the amount of flow going through

the well face decreases with the number of wormholes, effectively


limiting the maximum number of dominant radial wormholes
to between five and seven. Adding these results to the previous
model enabled scaleup from laboratory experiments to radial-flow
field conditions.
Fig. 5 is a summary of the 350F (177C) data projections for
a specific formation. In these tests, the calculated pore volume to
breakthrough (PVBT) is defined as the efficiency. Acetic acid was
less efficient than low-pH HEDTA at this temperature and flow
rate. The low-pH chelant fluid fulfills the criteria for ideal hightemperature stimulation fluids for carbonates because the corrosion rates can be adequately controlled11 to 400F. In addition, the
reduced dissolution rates allow fluid pumping to be controlled in
the range to achieve maximum efficient use of the fluid and avoid
fractures in the formation.
Stimulation of Sandstone. High-temperature sandstone acidizing
is challenging because of the very fast reaction rates and instability
of clays at these temperatures. Gdanski and Shuchert17 have shown
that essentially all clays are unstable in HCl above 300F (149C).
Techniques for solving the high-temperature acidizing dilemma
have included the borate retarded acid fluid developed by Thomas
et al.,18 as well as use of organic acid/hydrofluoric acid (HF)
blends.19,20
The ideal stimulation fluid would remove near-wellbore damage without depositing precipitates in the formation, thus eliminating well production declines caused by solids movement. The
ideal fluid would also exhibit very low corrosion rates at high
temperatures. Slurry reactor tests were conducted with HEDTA,
pH 4, as a solvent for several aluminosilicates at various temperatures. Fig. 6 shows the dissolution of kaolinite at 150C with this
solvent. Small amounts of aluminum and silicon were dissolved.
Fig. 7 shows initial rates for dissolution of various other aluminosilicates including a zeolite (analcime), three clays (kaolinite,

Fig. 2Projected optimal flow rates at specific core geometry for 20% Na3HEDTA at NDa=0.29.
November 2004 SPE Production & Facilities

191

Fig. 3Limestone cores after coreflood tests of four 20%


Na3HEDTA formulations at 250F (121C), 2 mL/min.

chlorite, and illite) and a feldspar (albite). The initial reaction rates
were all below 0.5 mmol/kg/min. For comparison, 9/1-mud acid
has an initial rate on kaolinite at 150F (65C) of approximately 10
mmol/kg/min. See Fig. 8 for a rate curve for this reaction. Therefore, we would not expect this chelating agent to damage acid
sensitive clays. A study by Hartman et al.13 demonstrated the
deleterious effects of mineral acids on acid-sensitive clays at much
lower temperatures than those seen in tests.
To determine the effects of using different types of stimulation
agents (chelate, HCl, or mud acid) on an undamaged core, three
fluids were tested at 350F (177C). Results of coreflood tests
using Berea sandstone (bulk analysis in Table 2) are in Table 3
and Figs. 9 and 10. Each test included two stages of 2% KCl to
determine initial permeability, a treatment stage [10 pore volumes
(PV) of acid] followed by two stages of potassium chloride (KCl)
to determine retained permeability. Table 3 shows the permeability
values at the end of the initial two KCl stages, at the end of the acid
stage, and at the end of each retained-permeability KCl stage. The
percentage of returned permeability is based on the final KCl
stage. The graphs show metals (Al, Si, and Ca) in solution during
the coreflood tests run at 5 mL/min through the 1-in.-diameter by
6-in.-long (25150 mm) sandstone core.
The chelating fluid Na3HEDTA stimulated the Berea core better than HCl and 9% HCl with 1% HF (9/1-mud acid) at 10 PV.

From the data in Table 3 as well as the permeability vs. time traces,
it is clear that the HCl and mud acid fluids reduced the average
permeability of the cores during the treatment phase. This is expected based on the work of Bryant and Butler,21 who found
damage to a majority of cores exposed to HCl at 176F (80C).
Table 4 shows the total Al, Si, and Ca conversions at the end
of the experiment for the fluids investigated in the study. Conversions are based on Al and Si present in the aluminosilicates (i.e.,
all minerals excluding quartz) in the core. The wt% amorphous
silica reported in the table is the amorphous silica still present in
the core at the end of the experiment. It was estimated as follows.
First, an average Al/Si ratio in the aluminosilicates was calculated
using XRD data. The ratio was approximately 0.56. Second, the Al
and Si concentrations in the effluent were integrated to obtain the
total amount of Al and Si brought into solution. Third, the expected amount of Si in the effluent was estimated from the Al/Si
ratio in the virgin core and the produced amount of Al. It was
assumed that no Al compounds precipitated in the core. Fourth, the
missing Si in the effluent was assumed to precipitate as hydrated
silica (SiO2.2H2O). The wt% reported in Table 4 is based on this
composition of amorphous silica. This is an estimated value, but it
serves as a good indicator for the performance of each fluid.
The data support a model wherein the permeability of the HClor HF-treated cores improves because amorphous silica precipitated by these acids is removed by KCl. The solubility of silica
increased dramatically as the KCl displaced the acid. The concentration of Si in KCl is in agreement with literature22 values on
solubility of silica in aqueous fluids at high temperatures. However, in a reservoir treatment, a post-flush after the acid stage
would move the precipitates deeper into the reservoir. This may be
recorded as an improvement in injectivity during the treatment as
the precipitate bank would move away from the wellbore. However, when the well is put back on production, the precipitates
would flow back and damage the near-wellbore area. Shaughnessy
and Kline8 noted damage after treatment of wells using HCl but
not after chelant treatments. If the reservoir pressure is not high,
the damage may be permanent, as these precipitates may never be
produced back. Furthermore, during an actual treatment, the fluids
would initially go to the high-permeability zones. These zones will
be damaged first, before the fluid diverts itself to the lowpermeability zones in the well. If the treatment is continued, successively all zones will be damaged. After the well is put back on
production only certain zones may clean up, and after this happens,
the damage from other zones may prove difficult to remove. Thus,
the remaining damage may never be adequately removed. Therefore, although we see a net improvement in permeability with
the HCl/HF fluids, it may not be realized in practice because a
large pressure differential may be required to flow back the pre-

Fig. 4Flow distribution between the wormholes and the well face based on radial flow simulations performed with the CFD code.
192

November 2004 SPE Production & Facilities

Fig. 5Efficiency of stimulation (PVBT) vs. pump rate for four acids at 350F (177C).

cipitated amorphous silica, and the reservoir may not be able to


provide the force.
Corrosion Tests. At high temperatures, one of the most important
considerations is corrosion control. While the art of corrosion inhibitor formulation has continued to improve control chemistries,
it still is difficult to protect many types of oilfield tubular steel at
temperatures >250F (123C). Organic acids such as acetic and
formic acid are also not acceptable above approximately 350F
(177C). Use of the higher-pH chelating agent formulations provides a great improvement in corrosion protection. Table 5 demonstrates that the pH 4 HEDTA formulations gave superior protection to both carbon steel and 13 Cr steel at 350F (177C)
compared with inhibited HCl or formic acid. The values for the 13
Cr steel would be unacceptable using the criterion of the authors
employers [0.05 lbm/ft2 (0.24 kg/m2) is the rejection value].
Based on these laboratory investigations, formulations containing HACA chemicals were used to stimulate several hightemperature oil and gas wells.
Field Treatments
Successful field treatments using chelating agent solvents that
have been conducted in sandstone and in carbonate formations
are described.

Treatment 1: High-Temperature Sandstone Formation. This


exploration well was drilled to a depth of 12,431 ft (3729 m)
through a western U.S.A. formation in 6.5-in. (165-mm) hole. The
well was perforated through 4-in. (115-mm) casing from 12,250 to
12,279 ft (36753684 m). The interval had an average permeability of 50 md, porosity of 12.3%, and a BHST of 375F (190C).
During the openhole-logging phase, numerous sidewall cores were
obtained to characterize the formation mineralogy. Table 6 shows
the results of XRD analyses. The table shows a sandstone formation with a very high clay content ranging from 7 to 29 wt%. The
predominant clays were illite, smectite, mica, and chlorite. In addition, a high concentration of sodium, potassium, and calcium
feldspars was also present (15 to 50 wt%). Acid-soluble minerals
were also identified (7 to 16 wt%) as mainly hematite, calcite, and
iron-dolomite.
The pay interval was drilled with a bentonite mud with a density of 9.4 to 9.7 lbm/gal (1.1 kg/L). It had suffered partial losses,
controlled with the addition of mica in the mud, and was not
producing any gas despite good prospects.
A mud acid treatment was initially suggested to remove the
near-wellbore damage and stimulate the well. However, there were
strong concerns regarding corrosion and the risk of inducing further damage as a result of contacting acid-sensitive clays with a
mineral acid treatment at such elevated temperatures. After highlighting such concerns, a pH 4 HACA fluid was proposed as the

Fig. 6Solubility of kaolinite clay in Na3HEDTA, pH 4 at 300F (149C).


November 2004 SPE Production & Facilities

193

Fig. 7Initial rates for dissolution of aluminosilicates in Na3HEDTA, pH 4, 150C.

basis for the treatment and the best means of stimulating while
minimizing adverse precipitation reactions. A hydraulic fracture
was also considered but was left as a contingency option because
of the comparatively high cost, coupled with somewhat uncertain
economic potential in a relatively new and underdeveloped field.
The results of initial solubility tests conducted at 180F (82C)
indicated that the formation samples were dissolved 12 to 15% in
15% HCl, 27 to 38% in half-strength mud acid (HSMA), and 2 to
6% in HACA pH 4. The tests involving HSMA exhibited significant precipitation after the 1-hour solubility test at this low temperature. Further solubility tests were conducted in an autoclave at
370F (188C). The data showed a significant increase in formation solubility with the HACA pH 4 solutions with no reprecipitation evident. Because no suitable core plugs were available, a
geochemical simulator23 was used to predict the radial skin generated after treatment with a traditional mud acid treatment. Fig.
11a shows the detrimental impact of a 5% HCl preflush followed
by a 6/0.5-mud acid and is consistent with the precipitation of
amorphous silica. The use of organic mud acid systems (Fig. 11b)
showed a mild stimulation during the formic acid preflush and a
significant increase in skin when the formation was treated with a
6:0.5 formic/ HF system. The conclusion here was that an HFbased treatment constituted a high risk of damage. Organic acids
such as formic or acetic acid were considered less of a risk in terms
of formation damage, but still pose a problem in terms of difficulties inhibiting these acids for prolonged periods at 375F
(190C). On the basis of laboratory and simulation studies as well
as field experience relating to the use of HF systems in hightemperature wells, a chelating agent solution was thought to provide the least risk and the best option to at least remove the calcium, magnesium, and iron carbonate minerals without inducing
damage through clay degradation and precipitated byproducts.
Treatment Design. The interval [12,250 to 12,279 ft (3675 to
3684 m)] was treated through 238-in. (60-mm) tubing set with a
packer at 12,080 ft (3624 m). A preflush of 3% KCl brine with
mutual solvent and a strong water-wetting surfactant, nitrified with
N2 to 50% quality, was pumped. A clay-control additive was in-

cluded to minimize dispersion of formation clays in the formation.


A second, nitrogen-energized fluid was pumped, incorporating 20
wt% of the HACA fluid with a mutual solvent, clay-control additive, corrosion inhibitor, and a strong surface-tension-reducing,
water-wetting surfactant. A small KCl brine spacer was pumped,
followed by the main treatment fluid, similar to the second fluid as
just described; but this time, the mutual solvent was omitted to
allow the surfactant to develop a stable foam. The intention was to
create 70% quality foam to promote full coverage of the interval.
The wellbore was displaced with nitrogen and flowed back immediately. The treatment was designed using reactive fluids placement software in conjunction with a step rate test, which provided
an estimate of the maximum pump rate achievable to stay at matrix rates.
Results. Fig. 12 shows the pressure and rate plot for the treatment. The surface pressure values show a constant increase during
the injection of the initial preflush, but there is a gradual pressure
breakback of 250 psi (1725 kPa) during the time the HACA fluid
is hitting the formation. The full effects of the chelant dissolution
of soluble materials in the formation are difficult to see with surface pressures, particularly as the treatments were energized. However, the 250-psi (1725-kPa) drop in surface pressure was observed
during a period when constant fluid and nitrogen rates of 1.7
bbl/min (0.27 m3/min) and 1800 ft3/min (50 m3/min) respectively
were achieved.
The main accomplishments are as follows:
The well was opened to flowback 15 minutes after shutdown.
The well brought back all fluids pumped and flowed at a wellhead
pressure of 1000 to 1200 psi (6900 to 13800 kPa) on a 1/32-in.
(0.8-mm) choke before conducting well tests.
This result represents a significant improvement because the
well was not producing gas when tested initially after completion.
Treatment 2: High-Temperature Oil Well. This well is located
in an offshore field in West Africa. The field is an elongated
anticline with some minor crustal faults. The reservoir thickness
ranges between 20 and 95 ft, with some small shale stringers

Fig. 89/1 mud acid used to dissolve kaolinite at 65C.


194

November 2004 SPE Production & Facilities

across the main reservoir body. The reservoir is deltaic sandstone


containing up to 15% carbonates across certain sublayers. The
reservoir BHST is 263F (128C).
This oil producer was drilled and completed in 1984 with a
total perforated interval of 105 ft (32 m). The deepest perforations
are at 7,765 ft (2367 m). Due to the low reservoir pressure, the well
was equipped with a gas lift system. Initial production rates averaged approximately 2,500 BOPD (490 m3/d).
Water production was first noticed in 1991 and has been slowly
increasing. As the water cut reached 30% in early 2003, with a
corresponding total liquid flow of 2,012 B/D (313 m3/d), a slickline intervention was planned to replace the gas lift valves. During
the drift run prior to the main slickline intervention, it was noticed
that there was a buildup of carbonate scales in the tubing. The scale
profile was such that a 2.79-in. (71-mm) gauge cutter could not
pass below 2,870 ft (875 m).
Because of concerns about tubular corrosion, ineffective stimulation of the formation, and handling and disposal of the spent
treating fluid, a pH 4 chelating agent (with corrosion inhibitor B)
fluid was chosen for the treatment. The goals included the following:
Use a fluid that during well cleanup could be routed directly
to the production separator (pH>5.5), thus avoiding the need for
neutralization and further pumping.
Efficiently remove carbonate scales from gas lift mandrels.
Do not damage the high-temperature sandstone formation
and metallic well internals.
Treatment Design. A coiled-tubing rig equipped with a highpressure rotating jetting tool was run into the hole. The first gas lift
mandrel was treated with chelating agent fluid (HACA pH 4)
containing a corrosion inhibitor by reciprocating the rotating jet
tool across the interval between 15 ft (4.5 m) above and 15 ft
below the gas lift mandrel. The treatment fluid was displaced with
clean filtered freshwater. The rest of the mandrels were treated in
sequence. A soak time of 4 hours was followed by resumption of
the gas lift to recover fluids while measuring pH and displace to
separator if acceptable.
Results. After allowing the chelating agent fluid to soak for 4
hours, the gas lift system was reopened, and the spent fluids were
recovered at the surface.
The pH of returned fluid was in the range between 6.2 and
7 (Fig. 13). Because it was above the minimum value, the fluid

was routed to the separator, allowing a significant savings in operating time (no need for neutralizing and pumping back to production facilities).
A 2.79-in. (71-mm) gauge cutter was run up to a 2.31-in.
(58-mm) nipple at 7,480 ft (2280 m) without finding any obstruction.
The post-treatment liquid rate was 2,528 B/D (402 m3/d)
with a water cut of 33.2%. The increase in oil production was 283
B/D (45 m3/d), indicating that the sandstone was not damaged. The
flow increased because of the removal of scale and possibly from
stimulation of the formation.

Fig. 9Metals during coreflood of Berea core with 9/1 mud acid,
350F (177C).

Fig. 10Metals during coreflood of Berea core with Na3HEDTA/


pH 4, 350F (177C).

November 2004 SPE Production & Facilities

Treatment 3: Carbonate Oil Wells with Gas in Dolomite


Layer. To improve the oil production and recovery from the field,
infill well drilling was accomplished based on the history match
black-oil models. As a result of the study, a water injector well
was recommended, with an oil gain estimated by the reservoir simulation of 4.1 million bbl (656000 m3) under a water
injection scheme.
The well described in this treatment was drilled as a water
injector in the crestal graben area, in order to displace the remaining oil trapped by the parallel faults toward the producer. However, there is a risk of the fault sealing, and the issue could be
clarified only after drilling the well. Therefore, it was completed
first as a producer to investigate the fault sealing surrounding the
well. The main concern in stimulating this well is the gas that
might evolve as a result of wormholing into the dolomite layers.
The BHST was 230F (110C).
The dolomite layers have entrapped gas that can evolve and
increase the gas/oil ratio (GOR). The operator could handle a GOR
of only 2500 ft3/bbl (440 m3/m3). For higher GORs, an alternative
means would be needed to handle the high production of gas on
surface, which would be costly. If we had stimulated the dolomite,
the GOR would have jumped to 5000 ft3/bbl (880 m3/m3). Fig. 14
shows the well log indicating the limestone and dolomite layers.
After completion, the well was not flowing and the formation
damage was suspected to be drilling mud, which had caused the
filtrate to invade into the matrix. The acid systems for this treatment were selected based on the lower reactivity of the chelants
with dolomites compared to limestone. The HACA chelants have
the capability to react with CaCO3, which is the main constituent
of the drilling mud that was used, similar to HCl. However, the

195

main reason for selecting the chelant is its enhanced reaction with
the limestone and limited reaction with the dolomite.
Treatment Design. The treatment was carefully designed with
the aid of a matrix stimulation design software package. The treatment consisted of using HACA pH 3, which was chosen for its
higher reaction rate than conventional chelants at this temperature
[230F (110C)], and the treatment was placed using coiled tubing.
A preflush of 10% mutual solvent mixed in water was used ahead
of the HACA solution to enhance injection. Water wetting of the
formation also will reduce surface tension during flowback. The
simulator predicted a radial penetration of 5 ft (1.5 m) of the
HACA into the critical matrix of the limestone. In addition, a
limited reaction of HACA with the dolomite layers was observed.
The simulator also predicted a skin evolution of 0 in the limestone
layer by the end of the treatment.
The average treating pressure and rate for the HACA treatment
was 1,200 psi (8280 kPa) and 0.35 bbl/min (0.056 m3/min). It was
observed that when the HACA reached the perforations, the pressure had declined by 400 psi (2760 kPa); therefore, the rate was
increased to ensure mechanical diversion across the zone. However, it was necessary to decrease the pumping rate because the
pressure was rebuilding as a result of pumping a lighter fluid into
the coiled tubing. After the well was displaced to diesel, the coiled
tubing was pulled 2,000 ft (600 m) above the perforated zone and
a lift was started with N2. As soon as the well was lifted with N2,
the coiled tubing was pulled out of hole.
Results. During treatment, a pressure drop was observed,
which demonstrates that the chelating agent did react with the
formation.
By stimulating the limestone and penetrating into the dolomite, we met the job objective. This was determined by production
on surface. It was estimated that if the dolomite streaks had been
stimulated and allowed to produce, the expected GOR could have
been 5000 ft3/bbl (880 m3/m3). With this kind of challenge, the
ideal stimulating fluid chosen for the job was the HACA, because
it has the feature of low reaction rate with the dolomite.
The HACA stimulation treatment had increased the oil production from 0 to 600 bbl/d (96 m3/d) with a GOR of only 1500 to
1700 ft3/bbl (264 to 299 m3/m3).
Conclusions
1. Laboratory tests and simulations have shown that HACA chelating agents can be used to stimulate carbonate and sandstone
formations at high temperatures. Because of reduced reaction
rates and corrosion rates, these fluids can be used to successfully
stimulate high-temperature hydrocarbon formations, where
strong mineral acids could damage the well tubulars and the
producing formations.
2. During field tests in sandstone and carbonate formations with
BHST to >365F (185C), fluids based on these chelating agents

196

have been used to successfully remove scale and stimulate the


high-temperature wells and to increase the flow rate of hydrocarbons. The ability to adjust the pH values to effect changes in
the rates of reactions allows the treatments to be individually
tailored to different well conditions. Fig. 15 shows a summary
of the improved well performance after the treatments.
3. Because of the near-neutral pH values of these materials, it was
possible to successfully dispose of the fluids in the platforms
treatment system without further chemical adjustments. These
factors, as well as the reduced toxicity of the chemicals, contribute to a reduced HSE footprint.
Nomenclature
C concentration, mol/L
d wormhole diameter, cm
Deff effective diffusion coefficient for reactants, cm2/s
Dp effective diffusion coefficient for products, cm2/s
Ea activation energy, kcal/mol
kr effective surface reaction rate constant, cm/s
ky apparent surface reaction rate constant, mol/cm2/s
Keq effective equilibrium constant
Kmt mass transfer coefficient, cm/s
Ky adsorption equilibrium constant, cm3/mol
l wormhole length, cm
m molal, moles/kg
NDa Damkhler number
q injection rate, cm3/min
T temperature, F (C)
stoichiometric ratio of reactants/products
overall dissolution rate constant, cm/s
kinematic viscosity, stokes
angular velocity, rad/s
Acknowledgments
The authors acknowledge the support and encouragement of
Schlumberger management for the writing and publication of this
paper. Dawn Alamia and Andre Roberson developed the coreflood
and dissolution-rate data; we also acknowledge the help of Don
Hill, who directed the corrosion studies, and Murtaza Ziauddin for
some of the coreflood data analyses.
References
1. Frenier W.W. et al.: Use of Highly Acid-Soluble Chelating Agents in
Well-Stimulation Services, paper SPE 63242 presented at the 2000
SPE Annual Technical Conference and Exhibition, Dallas, 14 October.
2. Frenier, W.W.: Novel Scale Removers Are Developed for Dissolving
Alkaline Earth Deposits, paper SPE 65027 presented at the 2001
SPE International Symposium on Oilfield Chemistry, Houston, 13
16 February.

November 2004 SPE Production & Facilities

3. Frenier, W., Fredd, C.N., and Chang, F.: Hydroxyaminocarboxylic


Acids Produce Superior Formulations for Matrix Stimulation of Carbonates, paper SPE 68924 presented at the 2001 SPE European Formation Damage Conference, The Hague, 2122 May.
4. Fredd, C.N. and Fogler, H.S.: The Influence of Transport and Reaction
on Wormhole Formation in Porous Media, AIChE J. (1998) 44, 1933.
5. Fredd, C.N. and Fogler, H.S.: The Influence of Chelating Agents on
the Kinetics of Calcite Dissolution, J. Col. Interface. Sci. (1998) 204, 187.
6. Fredd, C.N. and Fogler, H.S.: The Kinetics of Calcite Dissolution in
Acetic Acid Solutions, Chem. Eng. Sci. (1998) 22, 3863.
7. Coulter, G.R. and Jennings, A.R. Jr.: A Contemporary Approach to
Matrix Acidizing, paper SPE 38594 presented at the 1997 SPE Annual
Technical Conference and Exhibition, San Antonio, 58 October.
8. Shaughnessy, C.M. and Kline, W.E.: EDTA Removes Formation
Damage at Prudhoe Bay, JPT (October 1983) 1783.
9. Rhudy, J.S.: Removal of Mineral Scale From Reservoir Core by Scale
Dissolver, paper SPE 25161 presented at the 1993 SPE International
Symposium on Oilfield Chemistry, New Orleans, 25 March.
10. Huang, T., McElfresh, P.M., and Gabrysch, A.D.: Acid Removal of
Scale and Fines at High Temperatures, paper SPE 74678 presented at
the 2002 SPE Oilfield Scale Symposium, Aberdeen, 3031 January.
11. Frenier, W.W., Fredd, C.N., and Chang, F.: Hydroxyaminopolycarboxylic Acids Produce Superior Formulations for Matrix Stimulation
of Carbonates at High Temperatures, paper SPE 71696 at the 2001
SPE Annual Technical Conference and Exhibition, New Orleans, 30
September3 October.
12. Bazin, B. et al.: Improvement in the Characterization of the Acid
Wormholing by In-Situ X-ray CT Visualizations, paper SPE 31073
presented at the 1995 SPE International Symposium on Formation
Damage Control, Lafayette, Louisiana, 1415 February.
13. Hartman, R. et al.: Acid-Sensitive Aluminosilicates: Dissolution Kinetics and Fluid Selection for Matrix Stimulation Treatments, paper
SPE 82267 prepared for presentation at the 2003 SPE European Formation Damage Conference, The Hague, 1314 May.

14. Kung, M.S.: Flow and Reaction of Weak Acids in Carbonate Porous
Media, MS dissertation, U. of Michigan, Ann Arbor, Michigan (1998).
15. Lund, K. and Fogler, H.S.: AcidizationII. Dissolution of Calcite in
Hydrochloric Acid, Chem. Eng. Sci. (1975) 30, 825.
16. Fredd, C.N. and Fogler, H.S.: Optimum Conditions for Wormhole
Formation in Carbonate Porous Media: Influence of Transport and
Reaction, SPEJ (September 1999) 196.
17. Gdanski, R.D. and Shuchart, C.E.: Advanced Sandstone-Acidizing
Designs With Improved Radial Models, SPEPF (November 1998)
272.
18. Thomas, R.L., Crowe, C.W., and Simpson, B.E.: Effect of Chemical
Treatment Upon Formation Clays Is Revealed by Improved SEM Technique, paper SPE 6007 presented at the 1976 SPE Annual Technical
Conference and Exhibition, New Orleans, 36 October.
19. Van Domelen, M.S. and Jennings, A.R., Jr.: Alternate Acid Blends for
HT/HP Applications, paper SPE 30419 presented at the 1995 SPE
Offshore Europe Conference, Aberdeen, 58 September.
20. Shuchart, C.E.: Chemical Study of Organic-HF Blends Leads to Improved Fluids, paper SPE 37281 presented at the 1997 SPE International Symposium on Oilfield Chemistry, Houston, 1821 February.
21. Bryant, S.L. and Butler, D.C.: Formation Damage From Acid Treatments, SPEPE (November 1990) 455.
22. Iler, R.K.: The Chemistry of Silica, John Wiley and Sons, New York
City (1979) 76.
23. Ziauddin, M. et al.: The Use of a Virtual Chemistry Laboratory for
Design of Matrix Stimulation Treatment in the Heidrun Field, paper
SPE 78314 presented at the 2002 SPE European Petroleum Conference,
Aberdeen, 2931 October.
Wayne W. Frenier has recently retired from Schlumberger after
more than 30 years of research and development activities in
the energy industry. His most recent assignment was as a
Schlumberger Advisor in product development in Sugar Land,
Texas. He is the author of 23 U.S. patents, two books, and nu-

Fig. 11Geochemical simulator output showing the radial skin developed during (a) 5% HCl followed by a 6:0.5 mud acid
treatment, and (b) a 9% formic acid preflush followed by a 6:0.5 organic mud acid (6% formic acid: 0.5% hydrofluoric acid).
November 2004 SPE Production & Facilities

197

Fig. 12Treatment pressure responses during stimulation of a high-temperature sandstone with a chelating agent solution.
merous technical articles. Frenier holds an MS degree in chemistry from the U. of Chicago. Mark Brady is a Geomarket Technical Engineer for Schlumberger in the Arabian Gulf area,
where he is responsible for introduction of new technology and
development of stimulation services. He also has been responsible for research and development activities involving drilling
and stimulation services in North America and Europe. He is the
author of four U.S. patents and numerous technical articles.
Brady holds a PhD degree in chemistry from The Queens U.,
Belfast, U.K. Salah Al Harthy is currently an operations manager
for Schlumberger in Oman and is responsible for stimulation
services in the Arabian Gulf region. He has extensive experience with implementation of numerous well services, including

acidizing and cementing activities. He is an author of several


technical papers. Al Harthy holds a BS degree in materials science and engineering from the U. of Manchester, U.K. Roberto
Arangath is a Coiled Tubing Services Business Development
Manager for Schlumberger in the Europe and Africa regions.
Previously, he has held technical and engineering posts with
Schlumberger introducing and developing new well servicing
products. He is the author of several technical papers.
Araganth holds a PhD degree in chemical engineering from
the U. degli Studi di Trieste, Italy. Keng Seng Chan is a Principal
Production Engineer at Schlumberger Oilfield Services Technology Hub in Malaysia, responsible for production technical support for the Middle East and Asia regions. Previously, he has

Fig. 13pH vs. time plot for returns from Test Well 2.
198

November 2004 SPE Production & Facilities

Fig. 14Geologic prognosis of Treatment 3 Well.

Fig. 15Production improvements after HACA treatments.

held technical and engineering posts with Schlumberger, introducing and developing new well-servicing products. He is
the author of numerous technical papers. Chan holds a PhD
degree in chemical engineering from the U. of Florida. Nicolas
Flamant is a Senior Solutions Engineer in Schlumbergers Sugar
Land engineering center, where he is responsible for developing engineering models to support stimulation services. He is
the author of several technical papers. Flamant holds an MSc
degree from the Ecole Natl. Superieure des Mines de Nancy,
Parc de Saurupt, France. Mathew Samuel is an MEA client sup-

November 2004 SPE Production & Facilities

port laboratory and Schlumberger Well Services technology


manager based in Kuala Lumpur. Previously, he worked in
many positions, including Business Development Manager for
Stimulation (Middle East and Asia), Global BDM for nondamaging fluids technology, Technology Center manager, project
manager, and fluids specialist, among others. In these positions, he was responsible for the development and introduction of new services and products. He is the author of several
patents and technical articles. Samuel holds a PhD degree in
organic photochemistry.

199

You might also like