Você está na página 1de 45

Chapter 12: Chemical and Thermal Stimulation

When the formation will not produce at a rate equal to the expectations for the project, a stimulation
should be considered. There are two controls on the application of a stimulation treatment:

1 If permeability damage in the near wellbore area is blocking the production or injection of fluids,
and

2. If the permeability is low and the reservoir pressure is still high.

If either of the conditions are met and there are sufficient reserves remaining so that the increase in
production or injection will pay for the total cost of the stimulation, then the treatment can be justified.
Condition number one is the basic argument for a chemical treatment to remove damage. Condition
number two is the justification for fracturing (a few formations that cannot be fractured are chemically
treated to improve initial permeability). Current technology offers several methods of increasing pro-
duction from an existing wellbore, the two most common of which are hydraulic fracturing and matrix
treating. This chapter covers the basic principles involved in matrix treating and some of the more
common treatments. Matrix treating may involve any of a series of operations. They divide into chem-
ical and physical treatments.

Chemical stimulation is a collection of acid, gas, surfactant, reactant and solvent based treatments
that are intended to remove permeability damage or improve the natural matrix permeability through
nonfracturing methods. Physical treatments include perforating, perforation breakdown by nonprop-
pant fracturing (limited in this concept to linking perforations, small explosive detonations, and opera-
tions such as underreaming).

The type of treatment selected will depend on the type and extent of the permeability damage or the
permeability increase necessary to meet design criteria. The most efficient use of a chemical stimula-
tion treatment, as far as the well’s productivity is concerned, is to remove formation damage in a non-
fractured well or remove damage to the fracture proppant pack in a fractured well. The limit on matrix
stimulating an undamaged well can be seen by an examination of darcy law for radial flow through
beds in series.

where:

re = drainage radius, ft
r, = wellbore radius, ft
rl = radius of zone of improved permeability, ft
K, = permeability of improved zone, md
K2 = average reservoir permeability (unstimulated), md

12-1
Example 12.1
For a 15 md undamaged reservoir permeability on a 660 ft drainage area with a 6 in. wellbore, improv-
ing the permeability of the first 6 in. out from the wellbore face to a permeability of 1000 md would give
an average permeability of:

600
In-
0.5
Kavg =
1 1 0 1 660
-In- +-In-
1000 0.5 15 1

7.19
Kavg =
(0.001) (0.693) + (0.067) (6.49)

Kavg = 16.5 md

The productivity increase from a stimulation that produces this type of perm response would be:

0,
- = 16.5 (7.19)
= 1.21
Q, 15 (6.49)

If original productivity, Q,, were 100 BPD, then the improvement, Q2,would be 21 BPD or 21%.

Example 12.2
If the well had a zone of severe permeability damage of 0.005 md due to a mud cake, the effect of
even a 1/4 in. thick cake on the face of the formation would be:

600
In-
0.48
Kavg = = 0.84 md
1 0.5 1 660
In- + -In-
o.005 0.48 15 0.5
and the effect on productivity would be:

If Q, had been 100 BPD, the production through the damage, QP, would be 5.6 BPD or only a few per-
cent of the well’s potential. If the damage in this example were removed, the increase would be over
1000% of the damaged rate. Therefore; the rate of increase possible in treating a damaged well
depends on the amount of initial damage.

12-2
The basic equation shows that the increase in total flow rate at a given reservoir driving pressure is
about 10 to 20% for a well with undamaged permeability, even if the permeability of the first few inches
of the formation is raised almost to infinity. This does not mean that chemical stimulations to remove
damage are not useful: when damage removal is considered, the improvements in productivity of the
well may jump several times. The key to the use of chemical stimulations is whether or not the forma-
tion is damaged and if the damage can be removed.

Selecting a Candidate Well


Before any stimulation treatment is considered for a well, some kind of pretreatment analysis should
be run to determine the need for that stimulation and the cause of permeability reduction if the well is
damaged. Descriptions of several of these methods are given in the chapter on diagnostics. If a
buildup test, for example, indicates that the deliverability of a well is impaired by formation permeabil-
ity damage, then an analysis of well history and formation characteristics should be made to deter-
mine the type of permeability damage and its probable depth. If the damage can be removed by acid,
a stimulation to remove or bypass the effects of the damage may be chosen. Blindly trying an acid
treatment as a last-hope technique will usually result in large expense with little or no results. Poor
perforations, for example, will look like damage even on a buildup test but cannot be improved with
acidizing. A good evaluation is critical. Unfortunately, no test or set of tests can usually pinpoint the
type of damage; a deductive evaluation is usually required.

There are three basic types of chemical stimulation designs that encompass the vast majority of treat-
ments: wellbore cleanup, matrix treatment, and fracture acidizing. The choice between matrix treating
and fracturing may best be made after reviewing the type of damage and the materials available for
removal. To achieve economical damage removal and select the right treatment, two factors must be
considered: the type of the damage and the extent of the damage. Removing the damage, regardless
of the treatment method used, depends on removing the damage throughout all the wellbore contact.
If the damage covering only part of the wellbore is removed, the treatment results will reflect a much
smaller production increase. For this reason, almost all damage removal treatments need to be
diverted away from the undamaged area to the damaged area.

Evaluation of Stimulations
Beyond the usual need for quality control in application of a treatment, there is also a need to deter-
mine how well the treatment worked. Although there are a myriad of ways to evaluate a treatment,
what is needed most is an accurate assessment of what the well is capable of producing. Secondary
evaluations can then be made to optimize treatment techniques on a basis of money spent versus
results achieved. Difficulties encountered include variations in treatment action resulting from well-to-
well differences of the formation.

With the potential limits in mind, the needed data are:

1. an accurate assessment of what the well could make,

2. accurate knowledge of what the well is producing now,

3. the best idea possible of how the treatment affects the well’s performance on a step-wise basis.

The third element is often the most difficult to obtain. The most accepted way is use of a simulator that
combines treatment design with reservoir inflow and an economic module to forecast return on invest-
ment. The problem is that models usually treat formations as homogeneous; a necessity if the loca-
tion, size and parameters of the formation flow paths are unknown. The only way out of this ignorance
(classic lock of information) is to spend the time and money to find out what is needed, at least to
some level of satisfaction based on your bank account and your patience. Most companies are simply
not willing or able or determine every piece of needed information, hence, assumptions and approxi-
mations are what we use.

12-3
Selective Stimulation
The diversion of stimulation treatments or even diversion of heat during thermal operations requires
the use of diverting agents or processes to keep the fluid from going entirely into the zone of highest
permeability. Injection of any clean fluid into a formation is analogous with the flow of electricity; both
electricity and fluid flow follow the path of least resistance. Given equal wettability of the pores, the
fluid will preferentially flow through the larger pores of the formation where the effects of trapped fluid
attractions at the wall are at a minimum. In this type of a system, injection of any fluid into a reservoir
will result in fluid flow into and through the easiest penetrated sections. For many reasons, this type of
injection is not acceptable. To prevent its occurrence, some type of removable diverting process is
needed to evenly apply the treatment volume.

Diverting vs. Fluid Loss Control


The difference between diverting a stimulation treatment and establishing fluid loss control on a stim-
ulation treatment may be very similar in some cases and widely separated in others. Practically
speaking, diverting consists of allowing some of the treatment volume to penetrate the zone being
diverted and then slowly or rapidly diverting the remainder of the treatment volume to another zone.
The usual objective is to evenly treat the entire zone. Diverting is needed for permeability variations or
may be needed to seal off fractures. Fluid loss control is concerned with minimizing the volume of fluid
lost from a fracture treatment. Both diverting and fluid loss techniques may be used in either matrix
treating or fracture treating and in some cases, may use the same materials. For the purposes of this
discussion, fluid loss control will concern fracturing’ while diverting will focus on matrix treating.2

Candidates
Multiple pay zones require diverting since even thin separate zones are rarely of a similar permeabil-
Also, multiple treatments in the same single stimulation require the use of diverters for proper
separation or fluid loss controL5

The permeability variances are typically determined by grain size and sorting, mineral growth, bonding
of the grains and the presence of natural or induced fractures. When a formation is composed uni-
formly of a similar size sand, without severe mineral growth in the pores, the permeability to a single
saturating fluid is at a maximum. Poorly sorted formations, those having a wide range of different sand
sizes, are lower permeability because the smaller grains fill in the open spaces in the matrix around
the larger grains. Mineral growths such as clays in the pores also reduce the area open to flow.

Pressure within the formation can also dictate the route of fluid movement. Given two separate inter-
vals with contact to the same wellbore, one may have a much higher pressure than the other, resulting
in more difficulty in injecting into the higher pressure zone. If the higher pressure zone has been con-
trolled during drilling (no crossflow) and is controlled by the injection pressure during the treatment,
the major injection of fluid will be into the lower pressure zone. The treatment may be very successful
in removing damage, but only in the lower pressure zone. On a buildup test, the well will still appear to
be damaged. The higher pressure zone (possibly the most productive) will remain damaged and may
not contribute significantly to flow. A similar problem is realized when two zones are treated and one
has been previously stimulated. The previously stimulated zone, especially one in a producing well
that has been hydraulically fractured, is probably at a lower pressure because of depletion and will
accept fluid from the wellbore much more readily than an unstimulated zone.

Although the advantages of a more even stimulation treatment are readily apparent (e.g., more com-
plete processing and drainage of a pay), the advantages of either fluid loss or diverting in matrix treat-
ing can be overstated, particularly where large permeability variations exist. In normal operations, the
variance in permeability that can be easily treated by a diverting agent are about one order of magni-
tude of permeability lower than the highest permeability section of a well. This means that if the maxi-
mum permeability is 100 md, then the lowest permeability zone that can be effectively diverted to with
a large amount of fluid is about 10 md unless complete fluid loss control is established in the high
permeability zones. By using extremely efficient diverting or fluid loss materials (e.g., mechanical
diverting in a well with a good cement job and no fractures), it is possible to slowly drive fluid into

12-4
lower permeability sections. However, pay zones with permeabilities varying over two orders of mag-
nitude are rarely known to contribute from the zones of the lowest permeability before the highest per-
meability zones are depleted. This phenomenon has been proven in several areas of secondary
recovery operations where 20-year old waterflood patterns (40-acre spacing) have been drilled in an
in-field drilling operation and virgin pressures were found in the middle of the pattern in the lowest per-
meability pays. In this example, the permeability variation was from about 50 md to less than 1 md.
These wells had been repeatedly treated with diverted and undiverted stimulation treatments. These
procedures made little difference in the low permeability sections in the 40-acre spacing.

Diverter Operation
There are several types of diverters, techniques for diverter usage, and various kinds of mechanical
apparatus that control fluid flow or fluid pressure within a wellbore. The diverters are described by
whether the diverter is mechanical (a downhole tool) or chemical.

Limited Entry (Pressure Differential Diverting)


Limited entry is the act of placing only a few perforations in the pay. With the limited entry technique,
several pays may be treated in one operation.6i7 The process may be accomplished by using balls in
stages to seal off perforations after stages of a treatment or by using back pressure developed by flow
through a limited number of perforations. When using back pressure, the pays are perforated with the
high permeability zones receiving only a few perforations and lower permeability zones being shot
with several perforations. In this manner, each zone is forced to take a share of the total fluid. This
procedure has some special applications in single treating of several stringers of a formation: how-
ever, the limited number of perforations may create high perforation friction (800 to 1500 psi differen- -
tial pressures are often necessary for effective operation). In some treatments, the horsepower
necessary to overcome the perforation friction will be about 30 to 50% of the total horsepower used on
the job. For comparison of pressure drop, the flow rates through perforations at various pressures are

12-5
presented in Figure 12.1. Flow from the reservoir into the wellbore is also restricted in limited entry
completions and the well may appear to be damaged.8s9

A0 . DcP9
0.4367

(Scott et al.)

Figure 12.1: A nomograph method of determining


perforation induced friction.

Mechanical Diverting
If a rig is available or if coiled tubing can be used, packers and selective injection wash tools often
present the best method of matrix stimulation diversion, provided that the cement job is adequate to
isolate the annulus and the well is not fractured. Most mechanical diverting tools are severely limited
because they function only at the face of the formation.lo Fractures, high permeability streaks and
even channels in the cement will defeat mechanical attempts to divert fluid flow.

The mechanical diverters include selective injection packers, packer and plug sets, and other isolation
techniques that protect a section of the well from fluid pressure. The selective injection packers or
straddle cup packers isolate a section of perforations for and makes sure at least one of the perfora-
tions in the straddled interval is open. The first requirement of mechanical methods is that the cement
sheath must be in good condition so that fluids do not channel behind the pipe. The benefits of
mechanical methods include positive opening of perforations and, usually, effective breakdown of the
formation. The process is particularly suited for small volume treatments. The tools can be obtained
with almost any length of perforated tubing between the isolation cups or packers. As little as 2 in. or
as much as 50 ft or more may be built into the apparatus at the surface.

12-6
Disadvantages to the process, besides the required use of a rig, are leaks produced by wear on the
cups and seals during multiple sets of the packer and the density segregation of different fluids in the
tubing at very low pump rates. Wiper plugs to separate fluids are useful if there is enough room in the
perforated tubing to catch the plugs without restricting fluid flow.

Isolation packers are the most versatile mechanical diverting devices because they can be used and
reset numerous times on the same run. The packoff seal is provided by cup seals, Figure 12.2, or
packer elements, Figure 12.3. Cup packers allow faster sets and higher flow rates while hydraulic set
packers allow better seals inside the wellbore. Packer and plug sets, Figure 12.4, are a common
approach to separation of multiple zones where a fracture stimulation is planned.

1 nuRRlAFlR 1

E
I k v e valve is jayed in
ocked closed position. Fluid
umped into annulus will flow
rouah annular bvpass of
length of tool. I

t
-’
emain open M i l e running in.

(Baker)
Figure 12.2: Schematic and operational sequence of a packer using cup seals.

Through tubing bridge plugs are devices that can be run through a tubing string on a wireline and
opened in the casing below the tubing to provide a place to start a cement or sand plug.” A few
through tubing plugs are marginally effective without added sand or cement. The plugs are usually of
the petal basket, corrosion resistant fabric skin (resembles up-side down umbrellas) and bag type
designs. Plugs are set with wireline with from one to four or more runs required to set a plug. Cement
is then applied with a bailer. The limits on through tubing bridge plugs are dependent upon the design
and materials of construction. Models of the plug are available with vents and some, such as the bag
and fabric covered basket can be used in highly deviated wellbores. Inflatable plugs for both wireline
and coiled tubing application are currently available. These plugs will hold some pressure differential
but are a good base for sand or cement.

Ball sealers are a special method of controlling fluid flow.I2 These are small vinyl rubber covered balls
with a neoprene or syntactic foam center. The balls come in various densities and are designed to
plug a perforation that is taking fluid. By dropping several balls at once, a section of pay may be iso-

12-7
(Scott et al.)
Figure 12.3: Selective injection packer using
compression extended elements.

lated, allowing the next section treated without closing off part of the wellbore. The limitations, like the
mechanical devices, include the necessity of a good cement job. Additional requirements are round,
burr free perforation entry holes that will provide a good seal with the pipe. Also required is a little
understood quantity called “ball action”. “Ball action” refers to the ability of the ball to flow to and seat
on a perforation. The more fluid that a perforation is taking, the more likely the chance to seat a ball
sealer. For this reason, a well with a few perforations is easier to “ball o f f than a well with many perfo-
rations that are taking the same quantity of fluid.

To travel with a fluid, the ball must be kept in suspension with the part of the fluid that is moving
towards the Perforation. Density differences between the ball and the fluid are the major reason for
separation. Density differences of as little as 0.05 (gramskc) are sufficient at low liquid flow velocities
to allow the ball to sink into the rat hole or float up and remain under the packer. For best application,
the fluid should be in turbulent flow or the density of the ball should match the density of the fluid. A
flow rate sufficient to keep the ball suspended with the liquid in the tubing will probably be too slow to
keep the same ball suspended in the same liquid in the casing unless the density of ball and fluid are
the same. Although the flow rate will be the same, the velocity in the casing may be lower by a factor
of 6 to 10 than in the tubing. This reduction in velocity is often enough to allow the balls to separate
from the fluid. To overcome the loss mixing in the casing, the density of the ball should be as close to
neutral as possible (within 0.05 g/cc). An example would be using a 1.1 g/cc density ball in 15% HCI
(density = 1.145 g/cc).

The rate of fluid flow into the perforation controls the attraction of a ball. The inflow of fluid sets up a
vortex around the entrance hole, which aids in attracting solid particles such as ball sealers. Although
experiments have shown that a little as 0.05 bbl/min is necessary to seat a ball sealer, the chances of
attracting and seating the ball increase with the flow rate into the perforation.

12-8
(Clementz et al., 1982)
Figure 12.4: An example of an isolation technique
using a retrievable packer and a retriev-
able bridge plug.

The final factor in ball action is the number of shots per foot that are open to fluid flow. The best ball
action is seen when there are few holes in the section, forcing a high rate of fluid injection into the
open perforations. This maximizing of flow contrast increases the flow velocity into the perf and thus
the vortex around the perf. Most operators that use ball sealers successfully perforate at one shot
every other foot up to a limit of 4 shots per foot depending upon the permeability of the formation. At
higher permeabilities, usually over 100 md, the higher number of shots per foot can be used. At any
given permeability, as the number of shots per foot goes up, the success with balls goes down. Guide-
lines for use of ball sealers are shown in Figure 12.5.

A second velocity factor affecting balls is the downward component of velocity during pumping at high
rates and especially with balls that are significantly heavier than the treating fluid. If the end of the tub-
ing is spotted just above the perforations, a heavier ball will tend to retain the downward velocity of the
fluid and be carried into the rat hole. Spotting the tubing 40 to 100 ft above the top of the perforations
will allow the balls to spread out and slow to the velocity of the fluid in the casing.

Once the ball has been seated on the perforation there is little problem with the ball popping-off as
long as treatment pressure is toward the formation. Even high velocity flow immediately past the ball
will not dislodge the ball from the perforation.

The number of balls which are necessary to treat a pay zone depend upon the number of open perfo-
rations plus some excess number of balls. Estimates on number of balls needed for a successful oper-
ation range from 30% to 100% in excess of the number of perforations.

The balls are dropped in sets of one to a more than a dozen in most operations and are intended to
“ball-off’ a section of perfs that are taking fluid and divert the injection to other perforations. Injection
of ball sealers into the treating fluid is done on the high pressure side of the pump through ball injec-
tors. The best devices for ball injection are the positive displacement rod or plunger units. Star-wheel

12-9
Completion Length
Remarks on Use of Straddle Packers for Diverting
(perfed interval)
O-20ft Good petformance with spacing on 1 ft tor breakdown. Spacing on 1 to 3 ft for
acidizing or solvent inj.
20-100ft Spacing on 2 to 4 ft for breakdown or injection.
> 100 ft Spacing 4 to 10 ft for injection. Perf breakdown less effective because of large
spacing. Number of sets per trip should be less than 50. Cup or seal life is
shortened by burrs around the perforations.
Perfs per ft Remarks on Use of Neutral Density Ball Sealers
c1 to2 Best results from ball sealers. Use minimum 30% excess balls. Petformance
lessened by low perm, heavy balls, or low injection rate. Best results occur
when injection rate divided by the number of perforations is no more than
0.05 bbl/min/perf.
3 to 4 Marginal performance. Use 100% excess balls. Balls will not work well at this
shot density in low permeability formations.
>4 I
Ball sealers are not recommended.
Buoyant ball sealers with a density less than the carrier fluid are not recommended unless an interface
treatment is used.
Ball sealers heavier than 0.1 g/cc over the density of the carrier fluid are not recommended.
Figure 12.5: Selection Chart for Selective Injection PackerlBall Seal-

devices that depend on gravity feed of the ball to the slots in the wheel are less reliable and are not
practical for buoyant ball sealers.

Selection of the ball is usually confined to picking the density, although some size variation may also
be possible. The common ball sealer size is 7/8" diameter. Although this size is adequate for most
0.25" to 0.5" entrance holes produced by deep penetrating charges, the 7/8" balls are too small to be
used in the 0.75" to +0.9"holes produced by big hole charges with optimum casing/gun clearance.
Ball sealers also should not be used in most wells that have been reperforated, particularly if both per-
forating guns were zero degree phased. Ball sealers should not be used after a large volume proppant
frac because of possible casing hole enlargement by abrasion the p r ~ p p a n t . ' ~

In tests with tubing spotted above the perforations, the buoyant ball sealersl4 performed very poorly in
selectively shutting off perforations when used in a single fluid treating system with the packer set
immediately above the perforations. The buoyant balls (0.9 and 0.95) floated past perforations taking
as much as 0.1 BPM. Even when the ball entered the vortex created by the perforation flow, there was
only a small chance of seating the ball. Once the balls were seated, there was no problem keeping
them on the perforation as long as flow was maintained. Whenever the pump was stopped, the balls
immediately came off the perforations. As the hole diameter increases from the norm of 0.4"toward
0.6"or larger, the standard 0.75" or 0.87" balls may become wedged in the hole.

Heavier density balls are commonly used in fracturing operations where a very large amount of mixing
goes on and flow rates are extremely high. Ball densities from 1.1 to 1.3 are available. Some opera-
tors feel the heavier weights are necessary to prevent the balls from flowing back to the surface once
the treating pressure has been released and the well comes back on flow.

When the well flows back after a treatment with either lightweight ball sealers or neutral density ball
sealers, a ball catcher should be installed in the line to assist in catching the balls prior to flowing the
well fluids through the choke. A device of this type is shown in Figure 12.6.15

12-10
(Gabriel & Erbostoesser, SPE)
Figure 12.6: A ball sealer catcher to prevent choke blockage.

Chemical Divsrters
The chemical diverters include a wide range of products that are designed to work in one of three
methods; filling the perforation, sealing the perforation, or promoting diversion by means of viscosity.

The fill-type diverters, which include such materials as naphthalene, benzoic, graded salts, sand, and
other large particulate material are designed to completely fill the perforation tunnel and create a sec-
ond pack within the tunnel that will divert the flow of fluid into the formation. A sketch of fill diverter
operation is shown in Figure 12.7. The flow of fluid into the perf is not completely stopped, but the
treating pressure is raised by resistance of flowing through the fill in the perforation tunnel and the
treating fluid may be diverted to another perforation. Diversion with this type of a product is possible
even where the permeability of the diverter system is higher than the permeability of the formation,
since the area of the perforation has been reduced from the perforation tunnel wall area to the
entrance area of the perf. These types of diverters do have a lower limit on the formation permeability
at which they are effective. The permeability of the Yill” matrix must act as a restriction in the amount
of flow that a perforation will take. In very low permeability, non-naturally fractured formations (k c 10
md), these diverters will not work well since the permeability of the diverter pack will still be much
higher than the permeability of the formation. The exception is natural fractures that open wider at
treating pressure. The approximate effective permeability limits for various diverter packs and an esti-
mation of where they are effective is contained in Figure 12.8. Loading information for the number of
pounds per gallon and pounds per perf is also reported. This data was established using a laboratory
model of the flow system and takes into account the approximate permeability of the formation. Exam-
ples of the rate of fluid flow reduction for 100 mesh sand, granular salt, benzoic acid and naphthalene
are shown in Figure 12.9.16 Note that the final fluid flow after treating is not zero. The data shown in
the figure demonstrate that the larger particulate diverters all have a “pack permeability’’ that allows a
reduction in fluid flow into a zone but does not stop the flow entirely.

The second type of chemical diversion is forming a seal on the face of the formation. The diverters in
this category include organic resin dispersions and particulates up to 10 to 70 microns. The perme-
ability of the barrier created by these materials is nearly zero. These materials are most effective on
matrix fluid loss from either matrix or frac treatments but lose effectiveness when the formation is nat-
urally fractured. The products form a thin seal, Figure 12.1 0, by trapping of the micron-sized solid par-
ticles against the permeable face of the formation when the liquid enters the formation. The organic
resin slurry uses oil soluble, organic material in a water suspension to form a “mud-cake” or seal on
any surface where fluid injection Like any fluid leakoff control additive, the speed with
which resin products can control leakoff is dependent upon their concentration, Figure 12.1 1, and the
amount of leakoff. The barrier or dehydration cake (similar to a mud cake) formed is very thin but
impenetrable to flow toward the formation. A density of 1.04 g/cc allows easy placement without sepa-
ration problems. The drawback with the organic resin slurries are that they are concentration and sur-

12-11
Figure 12.7: Diverting with a fill type diverter. Although the diverter
pack has a higher permeability than the formation, a flow
resistance can still be created since the exposed area of
the diverter filled perforation is much smaller than the
exposed area of the open perforation.
As formation permeability decreases, the effect of fill type
diverting is lessened.

Diverter 1 Decomposition
o;;;/T I 2; 1 Soluble
In I Use
Concentration I

114 to 1
112 to 1
114 to 1
0.032 to 0.1 1 to 2.5
*. 112 to 5 gal fluid

.. per 1000 gal fluid


114 to 10 Ib per
1000 gal fluid
.* 10 to 30 Ib per
I I 1 I
1000 gal fluid
'Not suitable

I "Will not fill a perforation, acts only on or in matrix of formation


"'May be only marginally effective at best.

Figure 12-8: Guidelines for Diverter Usage in Non-Fractured Formations

face area dependent and do not work exceptionally well in fractures (where the surface area that must
be sealed is extremely large).

The diverter should not be readily soluble in the liquid being used for transport. Oil base diverters such
as most organic resin dispersions, for example, cannot be used in acid containing mutual solvents. In
these cases, an inert carrier fluid should be used for the diverter stage carrier. Clean up of the resin
barriers normally proceeds easily since almost all of the material is halted at the face of the formation
and the layer of resin will not hold pressure when the differential is from the formation toward the well-

12-12
(King & Hollingsworth, SPE 8400)

Figure 12.9: Flow reductions (in sand packs) possible with various diverters.

(King & Hollingsworth,SPE 8400)


Figure 12.10: Photograph of the barrier formed by dehydra-
tion (by leakoff) of organic resin slurry while
flowing water through a sand pack.

bore. Most of the organic resins are soluble in either condensate or crude oil and will be removed rel-
atively easily from the surface of the formation. A second type of small particulate solid that may be
used as a slurry contains inorganic material. The micron sized inorganic particulates will damage the
formation to some extent and their clean up depends on either removal by reverse pressure or slow
decomposition of the inorganic particle. Caution is advised in using this type of a product.

When naturally fractured formations are treated, a particle type diverter followed by or mixed with a fil-
ter cake type diverter should be considered if complete shutoff of the zone is necessary.

12-13
IGH PERMEABILITY RANGE

-54
H M

U
I

Y
U 1gaL/1,000 gal.
z
3
a
8
E
W

5 10 g a l R, #y) gaL 5 gal. I l , WO gal.


4
Y

U loo0
1 TOTAL VOLUME WATER T~ROUGHPACK (cc’s)
0 60 120
TOTAL VOLUME WATER THROUGH PACK (gallwrtlft.4

(King & Hollingsworth, SPE 8400)


Figure 12.11: The fluid leakoff control provided by organic resin slur-
ries is dependent on resin concentration as shown in the
pilot. Note that the concentration affects only the speed
of leakoff control, not the shutoff potential.

Viscous gels are a third method of diverting that can be effective in the wellbore, on the face of the for-
mation or in the formation. The viscous gels that are available at the current time, are usually poly-
mer19v20or surfactant gelled waters or foams.21 They make injection of the fluid into the zone more
difficult because of the viscous fluid’s resistance to flow. By consideration of the Darcy Law, one only
has to increase the viscosity of a fluid from-one to 100 centipoise, for example, to see that these fluids
radically reduce the amount of fluid lost to the formation. Drawbacks to the viscous polymer fluids are
the insoluble debris that are in most polymer systems and the residue remaining after breakdown.
Cleanup of deeply placed polymer systems can also be slow unless a properly functioning breaker is
used.

Foamed systems, where a gas phase is dispersed in a liquid phase, diverts by being difficult to flow
through the f~rmation.~’-~’ Gas-in-water foam is an emulsion and are more viscous than the unvis-
cosified normal treating fluids. At least part of their resistance to flow comes from a process governing
droplet or bubble deformation, known as the Jamin e f f e ~ t . 2This
~ effect describes the difficulty in forc-
ing the droplet or bubble to deform sufficiently to flow through the pore, Figure 12.12.

It is not unusual to see a combination of two or more of the described systems or products to achieve
better fluid loss control, especially where formation permeability variances are large.2426

The carrier fluid for a chemical diverter is of prime i m p ~ r t a n c eThe


. ~ ~carrier fluid must be inert to the
product and should not modify the size of the product. In most stimulations the diverter or fluid loss
additive must be carried by the treating fluid. This adds another limit to the selection of the diverter.

interface Treatments
The interface technique is a procedure for directing placement of fluids or divert er^.'^ This process
uses a heavy fluid and a lighter, normally immiscible fluid. A brief list of the fluid densities available for
this work is shown in Figure 12.14.

12-14
Water w-
-q-
HIGH PERMEABlLlTY 4

Water

LOW PERMEABILITY

Water +

(Penny)
Figure 12.12: Mechanism of fluid loss control during an immiscible phase injection

Diverters or ball sealers with a density between the two fluids may also be used. If a diverter is used
and the density of the two fluids are selected correctly, the diverter must be localized at the interface
of the fluids. The treatment can be used with or without tubing in the well when immiscible fluids are
used, although the control of the interface is made easier if the tubing is spotted through the zone to
be treated. When the tubing extends through the zone, the position of the interface in the annulus can
be controlled by injecting heavier fluid into the tubing and lighter fluid into the annulus.

The position of the interface may be monitored by the use of a gamma ray tool in the tubing if the
annular fluid (upper fluid) is tagged with a tracer that is not soluble in the second fluid.

The position of the interface does not identify the injection site of the fluid unless injection rates into
both annulus and tubing are equal and there is only one zone taking fluid. If only one of the fluids is
being injected, and the other side shut in, the position of the interface only signifies that injection into
the reservoir is taking place either above or below the interface depending on which fluid (annular or
tubing) is being injected.

DiverterlFluid Loss Control Stages


Placement of the materials within the treating volumes of a stimulation treatment requires information
on the extent of leakoff into the formation at vmious times. In an acid fracturing treatment, the use of
fluid loss additives must occur early and frequently during the job for an optimum operation. Acid
increases the permeability of the formation by reaction and thus increases both the number and the
severity of fluid leakoff sites. To produce deeply penetrating acid fractures, initial acid leakoff must be
stopped and all subsequent deterioration of the fluid loss additive on those sites must be minimized.
Acid fracturing requires more fluid loss control than a fracturing treatment or other type of non acid
stimulation. In stimulations that use acid, the fluid loss additive is added on a regular basis either in
closely spaced stages or continuously in small quantities.

Placement of the stages is often done without much formation information and normally accounts for a
waste of a large amount of the treating volume, especially in the later part of the job. The optimum
place for a stage in an acid fracturing job is when the bottom hole injection pressure begins to decline
during a steady rate treatment. In fracturing, this pressure decline is most often caused by a reduction
in back pressure and signals a reduction in the amount of fluid traveling down the length of a fracture.
This indicates leakoff (usually severe) through the walls of the fracture.

In matrix acidizing, the diverter stage should be pumped when the acid causes a reduction in injection
pressure by increasing the permeability (or removing damage) in the zone into which it is flowing. This
pressure reduction may often be seen on the surface pressure recorder as a slow decrease followed
by a stable pressure when the damage has been removed. The diverter stage should be pumped
when the pressure begins to decline; this will allow the acid in the tubing to continue going into the

12-15
Pressure
Density
Fluid Unit Depth

Figure 12.14: Gradients and Densities

treated zone to finish the cleaning job and then divert the next volume of acid to a new zone. If the
diverter is properly selected and placed, the surface pressure should rise slightly as acid is injected in
a damaged zone then decline as a new interval is cleaned up.

Practically, and especially in deep wells, this method of treating is difficult due to the large volume of
acid in the tubing and the slow rate of displacement. The best method of designing diverter stages for

12-16
these deeper wells is to examine treating reports and pressure charts of jobs on offset wells or wells
with similar damage conditions. Select volumes for acid stages which approximate the quantity of acid
needed to reduce the injection pressure. The quantity of diverter and the volume of diverter stage may
be selected by examining treating reports to see if the diverter stages were effective. To place the
stages in the exact spot in the treatment where fluid loss control is needed would require advance
knowledge of the exact leakoff behavior of the formation. This information is not available for any but
the most homogeneous formations. The position of the first stage in the job may be approximated by
examining the treating records of an offset well. In wells with damage, there may not be enough con-
sistency between wells to make this process useful.

If the well is shallow and the injection rate is high or the treatment is fairly large, the surface pressure
can be monitored to know when to start a stage. An illustration of this technique is shown in
Figure 12.1 5. This example dramatically shows the fluctuation of injection pressure and fluid loss in a
fracturing well. It is interesting to note that the 100 mesh sand used as a fluid loss control in this exam-
ple has a pack permeability of 200 to 800 md, yet is able to reduce leakoff through open natural frac-
tures in a formation with a matrix permeability of 1 to about 15 md. The effective contrast is between
the fracture flow capacity and the sand pack, not between the sand pack and the formation.

\ f h l f l l Y G F01YA1101: -ACID
I ---?AD
I
1
I
\

E A U PAD STAGE
CONSISTS Of 1000
CA1 WITW 1 ))( f l l f
I f S H SAND EACH ACID
SIACE. 1 5 0 0 SAL

I
~ ~ ; ; ; o I o ; 15 r ao ' a s' '
40 '41 ' ' I
TIYI. IlNUltS
Y

(Coulter et al., SPE)


Figure 12.15: Plot of bottom-hole pressure change during constant pump rate
treatment of Strawn Reef Formation.

HCllHF Treatment Diversion


HCI/HF treatments pose a special problem for diverting since in most cases, an HCI preflush is
needed prior to the HCI/HF acid injection. If the HCI is not ahead of the HCVHF acid in each zone it
enters, damage from reaction of HF with calcium carbonate (results in CaF2, a precipitate) or forma-
tion waters may occur. The easiest and most effective way to divert HCVHF jobs is to treat each stage
as a separate treatment. The HCI and HCVHF acids are followed by a stage of 2% NH4CIwater with a
fluid loss additive. Salt should not be used as a diverter in HCI/HF acid, because it will create a
sodium fluoride precipitate.

Recommendations
The suggested loading rate of diverter materials for matrix operations were shown previously. These
loading rates are only a guide; actual use may have to be tailored to achieve optimum performance.
Very large (over 1 Ib/gal) loadings of flake diverters should be avoided due to potential bridging prob-
lems in the tubing.

The common diverters are listed in the Appendix 12.A.

12-17
Cleanup
Regardless of the type of material that is injected as a fluid loss or diversion additive, very careful
thought must be given to cleaning the material from the formation or the fracture pack.28 If the material
cannot be removed easily during the course of the well clean up, much or all of the advantages pro-
duced by the stimulation will be completely lost. Common solvents are available for almost all of the
commercially available diverters and fluid loss additives used in well stimulation.

Cleanout of diverters from the well is often best accomplished by backflow at high pressure differen-
tials or extended production. In tests on drilling mud and diverters on sandstone cores it has been
determined that approximately 80% of the initial permeability is usually attained when the core was
backflowed at a pressure differential at least equal to the pressure used to place the mud or diverter.
Final cleanup may depend on the diverter used and the type of overflush. In most cases, the overflush
should be a solvent for the diverter.

The worst problem encountered in removal of fluid loss control additives is in the drilling additives.
Lost circulation material (LCM) has been a historic problem because of drillers’ insistence on using
cheap materials to control fluid loss sites. In non pay sections the removal of these materials is of little
importance; however, when the fluid loss zone occurs in a pay, additives such as paper, leather, grain,
plastic, or other insoluble materials absolutely should not be used. Techniques for control of the most
severe fluid loss zones are a~ailable.~’

Packoff Techniques
Packoff techniques are normally used to isolate producing perforations from a treatment on other per-
forations where bridge plugs and other mechanical devices are impractical. These packoff techniques
include particulate fill and gelled plugs.

Crosslinked gelled plugs of several thousand centipoise viscosity are typically high concentration
polymer pills at a loading rate of 100 Ibs or more of polymer per 1000 gal of water. The polymer is
pumped with a time or temperature delay crosslinker that will render a solid plug. The polymer system
can be stabilized for long time stability. Problems with the gelled plugs have been with failure of break-
ing systems and cleanup of residue. Bacterial degradation may also be a problem.

Technology of Bridging
The size range of particles necessary for control of fluid entry into a formation will depend on the pore
size range or fracture size in the formation. Data from gravel packing studies has shown that spheres
can be bridged on an opening twice their diameter.30 In other tests, particles as small as 1/3 the open-
ing diameter may rapidly form bridges on an opening. In large particle concentrations such as sand
slurries, bridging may even occur with particles as small as 1/6 the opening diameter. In general, the
larger the particle, the faster the bridge is formed and the more stable it is. Use of polymer gels to
transport the particles or a second bridging material may also have an effect on bridging3’ As particle
size decreases in relationship to opening size, the ease of forming a bridge also decreases. More
rapid bridging can usually be achieved when a wide range of sizes are used. The larger particles will
bridge off on the opening and the smaller particles will bridge off on the remaining openings between
the large particles and the original opening. This analogy works for pores or natural fractures.

Depth of Diversion
Once the fluid enters the formation, regardless of the point of entry, the path of the fluid will be the
path of least resistance. The natural tendency will be to flow toward the region of lowest pressure
through the most permeable path available. For this reason, the near wellbore is most affected by the
action of the diverter. Without barriers to flow between the high perm and lower perm areas, most
diversion attempts will simply open new channels of permeability from the affected perforations to the
zone of highest perm.

12-18
Chemical Stimulation Techniques
Acidizing
Of the four most widely used acids, hydrochloric acid (also referred to as muriatic) is the most impor-
tant due to its high carbonate dissolving capacity and low cost.

Hydrochloric acid reacts with carbonate formations to form water, carbon dioxide gas, and calcium
chloride, as shown in the following schematic.

HCI Acid + CaC03 Limestone -+w2° Water + CO2 Carbon Dioxide + CaC12 Calcium Chloride
At bottomhole pressures, the CO2 produced is dissolved in water and remains trapped until the pres-
sure is lowered.

One thousand gallons of 15% HCI acid will dissolve 1840 Ibs of limestone (10 cubic ft3 if porosity =
O).32 The products formed are 2000 Ibs of calcium chloride, 8010 Ibs (6600 standard ft3) of carbon
dioxide gas, and 333 Ibs (40 gallons) of water. The total volume of water remaining after complete
spending of the acid would contain 15% (by weight) calcium chloride. Limestone dissolved by 4 m3 of
15% HCI would be 884 kg with 961 kg calcium chloride, 187 m3 CO2 at standard conditions and
160 kg (0.15 m3) water produced.

The reaction for dolomite is similar to the reaction for carbonate, but slower at temperatures of under
200°F (93°C).

HC1+ MgCa (CO,) , 'H,'O + CO, + CaCl, + MgCl,


Acid Dolomite Water Carbon Calcium Magnesium
Dioxide Chloride Chloride
Gas

The reaction of 1000 gallons of 15% hydrochloric on dolomite would consume 1700 Ibs of rock (9.6 ft3
if porosity = 0) and produce 1000 Ibs calcium chloride, 870 Ibs magnesium chloride, 6600 standard ft3
of carbon dioxide, and 40 gallons of water. Dolomite dissolved by 4 m3 of 15% HCI would be 817 kg
with 480 kg calcium chloride, 41 8 kg magnesium chloride, 187 m3 CO2 at standard conditions, and
160 kg (0.15 m3) water produced.

As better inhibitors have been developed, the higher concentrations of hydrochloric acid have come
into greater use. The use of 28% HCI for fracturing provides about twice the dissolving capacity at less
than twice the cost of 15% HCI. The 28% HCI is normally used for acid fracturing in carbonates.

Although HCI reacts readily with the calcium and magnesium carbonates, other acid reactions occur
during the treatment that can create damage if not anticipated. Iron, contained in such forms as rust,
pyrite, pyrrhotite, siderite, magnetite, and hematite, is a prime problem reactant with HCI because of
the possibility of iron hydroxide precipitation from spent acid. Normally, only magnetite and hematite
(+3 valence states) are troublesome since they precipitate at a pH of 2.2; however, the other iron com-
pounds (+2 valence states) can also precipitate if the acid spends completely (pH 2 7) Before selec-
tion of an iron sequestering agent, a good understanding of the iron precipitation problem should be
acquired. Other possible reactants with HCI include iron sulfide (a corrosion product present on tubu-
lars in sour gas areas) and HCI-soluble scales and clays.

12-19
HWHF: HCI/HF is a mixture of hydrochloric and hydrofluoric acids. This combination is used exclu-
sively in sand reservoirs that contain very little calcium. Acids containing hydrofluoric are not used in
calcium containing reservoirs since one of the reaction products, calcium fluoride, is a precipitate.

HF + CaCO, + CaF, + H,O + CO,


Hydrofluoric Calcium Calcium Water Carbon
Acid Carbonate Fluoride Dioxide

The calcium fluoride is an insoluble product and cannot be removed with normal treating processes.
To minimize harmful by-products when designing an HCI/HF acid stimulation, it is necessary to use a
hydrochloric acid preflush and a gas, HCI acid, ammonium chloride water, or hydrocarbon afterflush.
In dry gas reservoirs or elsewhere that oil would create relative permeability effects, the HCI/HF acid
should be followed by ammonium chloride water or gas. HCI/HF acid treatments should not be pre-
ceded or followed by waters weighted with sodium chloride, calcium chloride, or potassium chloride.
These salts, although relatively inert, can react with any unspent hydrofluoric acid to form insoluble
fluoride precipitates. A water solution of ammonium chloride can be used as a preflush or afterflush for
HCVHF acids without creating insoluble by-products.

In formations where the calcium content is low (i.e., less than 1O%), a preflush of hydrochloric acid is
required to remove the calcium which is in the pore throats. In formations of over 15% calcium com-
pounds, HCI/HF acids should not normally be used.

The reaction schematic for HF on an idealized clay is shown below.

HF + A1,Si,'0,6(OH), + H,Si'F, + AlF, + H,'O


Hydrofluoric "Clay" Hexafluro Aluminum Water
Acid Silicic Fluoride
Acid

The most common concentration of HCVHF is 12% HCI and 3% HF. In formations that can be dam-
aged by HF by-products, half strength HCVHF, 6% HCI and 1.5% HF, is popular. HF acid spends rap-
idly on clays and silts. In formations containing 10% clays, the HF acid will probably penetrate no
further than about 8 in. from the wellbore and removes approximately 120 Ib (54.5 kg) of clay per
1000 gallons of the HCI/HF mix before being completely spent. If damage due to clay or to completion
fluids extends deeper than the live HCI/HF acid can reach, it is advisable to consider an in situ HF acid
or a fracture treatment with proppant and water or oil-base fluid. HCI/HF is also very
useful for removing drilling mud cakes and dispersing mud.

HF Acid Spending Rates


Recently, work by Gdanski has illustrated potential problems with HCI/HF by-products when zeolites
(a reactive mineral) and a few other, highly reactive clays or minerals are encountered. In the work,
Gdanski shows that aluminum from the minerals precipitates as the acid spends (pH rises). The pre-
cipitation seems to be affected by both the aluminum concentration and the pH: themselves a function
of HF acid content and total acid content, respectively. To minimize potential precipitation and forma-
tion damage, the HF content is reduced and HCL is increased when acidizing some formations. Typi-
cal of the acids is an 7-1/2 to 10% HCL with a 1% HF.

Acetic and Formic: Organic acids are used in stimulations where their slower reaction time and ease
of inhibition are required. The acids most frequently used are formic and acetic. The carbonate dis-
solving capacity of the 10% organic acids is regulated by a reaction equilibrium between the reac-
tants, the product CO2 "gas" and the pressure. At pressures of over 500 psi, and up to about 160°F,
10% acetic will dissolve approximately 420 Ibs of calcium carbonate per thousand gallons
(202 kg/4 m3), while the 10% formic will dissolve roughly 750 Ibs of calcium carbonate per

12-20
1000 gallons (360 kg/m3). At atmospheric pressure, the 10% acetic will consume 740 Ibs of calcium
carbonate per thousand gallons (355 kg/4 m3) and the 10% formic will dissolve 940 Ibs of calcium car-
bonate per 1000 gallons (451 kg/4 m3). Since the spending rate of the organic acid and total material
consumed is controlled by an equilibrium with temperature and pressure as two of its controls, live
acid will usually be returned to the surface following a treatment with organic acid (especially at tem-
peratures below 160°F).

On a basis of cost per Ib of carbonate dissolved, the acetic acid is roughly five times the cost of the
hydrochloric, while the formic acid is about three times the cost of HCI acids. In high-temperature for-
mations, the cost factor between organics and HCI acid narrows due to the cost of special inhibitors
necessary for HCI acids at high temperatures.

Formic and acetic acids are not normally used at over 15% strength due to solubility limits of calcium
formate or calcium acetate, the chief byproduct. The reaction schematic of formic acid is shown below.

HCOOH + CaCO, + Ca(COOH), + CO, + 'H,'O


Formic Limestone Calcium Carbon Water
Acid Formate Dioxide

The reaction of acetic acid is similar except calcium acetate is formed instead of calcium formate.

Although the organic acids may be used by themselves for stimulation at high temperatures, it is often
advantageous to use the acids in a mixture with hydrochloric acid.

Solvents
Solvents cover a broad range of materials that dissolve and disperse deposits and damage problems
in the well. The most common solvent is fresh or brine water, used to remove salt, or as a base fluid to
carry surfactants, alcohols, mutual solvents and other products. Alcohols are a special class of sol-
vents since they have solubility in both oil and water. Hydrocarbon solvents are also used with regular-
ity.35-38These materials include crude oil and condensate, plus refined oils such as diesel, kerosene,
xylene and toluene. The reasons for the use of solvents are that acid has little or no effect on may
damaging deposits. Selecting a solvent usually requires some testing with the damage deposit. A few
selected organic solvents are shown in Figure 12.1 6.

Fluid Used to Remove

Methyl aiconoI
Diesel
Kerosene
Toluene
Xylene (meta)
Xylene (para)
Xylene (ortho)
Xylene bottoms
Naphtha
Oil
Gasoline
~

Figure 12.16: Common Organic Solvents Used in Treating

12-21
Gases
Although not usually considered as treating fluids, carbon dioxide gas and nitrogen gas are being
increasing used as additives to stimulation treatments and by themselves to help remove damage.
Gasses in an injected fluid provide assistance in fluid recovery by expansion in the reservoir when the
surface treating pressure is released. The expanding gases not only propel liquids from the reservoir,
they also provide a gas lift to produce the fluids up the tubing. Gas injection can also help restore gas
saturation in water blocked gas wells by driving the water out of the pores of the formation. A special
use of carbon dioxide gas has been in CO2 “huff and puff treatments where the CO2 is injected into a
heavy oil zone to swell the oil and reduce oil viscosity as well as providing driving energy to produce
the oil.

Surfactants
Hundreds of chemicals are available as surfactants, and each has a use in removing or preventing a
damage problem. Surfactants are used in concentrations of 5 ppm to 1% or more in a carrier fluid that
may be water, acid, or oil. Their properties are such that they congregate at the high energy interface
or surface and influence the formation and stability of emulsions, foams, sludges, surface tension,
particle suspension, surface wetting, scale growth, paraffin precipitation, and film interactions. Surfac-
tants are common in most treatments but tend to be overused. Surfactants should be selected on the
basis of carefully run, lab or field tests and only the essential surfactants used in any job. Surfactants
can react with each other when several types are mixed.

Reactants
Reactants are a group of materials that enter into chemical reactants with materials in the well but are
not acids. These materials include bactericides, oxidizers such as bleach and chlorine dioxide, chelat-
ing agents, and others. They usually have a limited number of specific purposes and must be matched
very carefully with well conditions. Their use in the oilfield is steadily increasing.

Treatment Types
A wellbore cleanup treatment is used to remove cement residue, drilling mud particles, scale, and per-
foration debris which cause injection face damage. Solvents such as diesel, xylene, kerosene and
alcohols are common as well as the mineral and organic acids. The acids used in these stimulations
are normally hydrochloric acid, HCI-HF acid (a mixture of hydrochloric and hydrofluoric acids) and,
less frequently, organic acids such as acetic and formic. The concentration of these acids for the
cleanup treatment varies from 3% to 15%. The wellbore cleanup treatment, using either acid or sol-
vent, is administered in the form of a soak or a slow injection, and the volume of the treatment is only
slightly larger than the tubular volume across the treatment zone. The solvents such as xylene are
often useful if there is paraffin, asphaltene or oil sludge damage. Alcohols and mutual solvents are
used to break emulsions, strip oil coatings, remove water blocks and alter wettability.

After the stimulation or cleanout has been accomplished, the solvent or acid should be produced from
the wellbore by either swabbing, pumping, displacement with nitrogen, or allowing the reservoir fluids
to backflow the material. Only in injection wells which handle large (>500 bbl/day) volumes of water
should the treatment be displaced into the formation without attempting to backflow. Note: Some wells
cannot be backflowed and the treating fluids must be injected into the reservoir and displaced without
return.

Matrix treating is normally used to increase the permeability of the formation immediately surrounding
the wellbore or to remove permeability damage that is beyond the injection face. Matrix treating with
solvents is useful for removing almost any hydrocarbon base damage in the pores. The reaction rate
of solvents in the matrix is usually slow and is limited by the amount of solvent that can be injected
into the pores to contact the damage. In matrix acidizing, the acid is also often used as a carrier to
transport surfactants such as clay stabilizers or emulsion breakers. The minimum volume of acid
which should be used in a sandstone matrix acidizing treatment will depend upon the type and depth

12-22
of damage, the permeability of the formation and the general response of the acid on the formation.
Large acid volumes on high permeability sandstones can be beneficial.39

Overflush volumes on low permeability zones should be only to the perforations. If large volumes of
acid or overflush are used on low permeability zones, the formation may be damaged since acid reac-
tion products may come out of solution before the acid leaves the formation.

Damage Removal
During pumping in matrix acidizing, as in any treating procedure, useful data on treating effectiveness
can be deduced from surface pressure recorders. It should be remembered with any surface pressure,
especially during changeovers from one fluid to another, that the pressure shown on the recorder
reflects the injection pressure of the fluid currently at the perforations minus the hydrostatic pressure.
The friction pressure, if any, increases the surface pressure. Any change in density of fluid in the tub-
ing will change the surface pressure at constant injection rate. Thus, any change that an injected fluid
makes on the formation will not be indicated by the pressure recorder until the newly injected fluid
reaches the perforations, usually several minutes after the pumping begins. With most acid systems in
permeability damage removal treatments, the pressure recorder may register a small rise (compared
to water or preflush injection) as the acid enters the formation followed by a gradual decrease to some
stabilized pressure. The timing and the sharpness of the surface pressure drop is determined by depth
and amount of the damage and the reactiveness of the acid toward that damage. For example, cal-
cium carbonate scale damage can be removed relatively quickly while particle stabilized emulsions,
deep clay damage and thick scale deposits are removed very slowly.

Viscous stable emulsions and water blocks clean up very slowly and may require soaking or a slow
injection with the mutual solvent/acid system. Emulsion breaking in the pores of the formation is vastly
different from a simple beaker demonstration showing breaking of the emulsion by surfactants. Break-
ing emulsions (and removing water blocks) requires contacting each droplet of the emulsion with the
treating fluid. A long cleanup time, or even repeat treatments in the case of a large amount of emul-
sion, is normal. For more information, see the description of on emulsions in the chapter on formation
damage.

If a continuous pressure rise is noted during the injection of a clean fluid, this may be an indication of
detrimental acid reactions occurring in the formation. These acid reactions may include formation of
sludges, release of migrating particles, clay swelling, creation of emulsions, or other reactions. If the
pressure climbs sharply and continuously for several minutes, pumping should be stopped and the
well backflowed. Samples of this backflow should be caught for analyses of iron content, acid
strength, and presence of emulsions and solids.40 If the formation cannot be broken down, the loca-
tion of the perforations should be checked. If the perforations appear to be at the right depth, reperfo-
rating is suggested.

In most instances of acid creating damage in oil wells, emulsions or sludges are usually at fault. The
sludge or emulsions formed between acid and a few crude oils are worst when iron and asphaltenes
are present and are very viscous and stable.41142

In gas wells, the most likely damage from an acid treatment will be water blocks created by the spent
acid. If water blocks are known to form in the subject formation, either mutual solvents or alcohols
should be added to the entire treatment.

If the injection pressure climbs rapidly during the treatment and the well will not backflow, formation
damage, or mechanical problems may have occurred. Checks of the mechanical equipment (including
packer location) and reactivities of oil, acid (with additives), and formation water should be made.

If the surface pressure decreases rapidly during the treatment or the well goes on LLvacuum”(i.e.,
takes fluid rapidly with hydrostatic pressure only), (1) matrix permeability is being increased (or dam-
age removed), (2) natural fractures are being enlarged, or (3) the formation has been fractured. This

12-23
behavior is very common on low pressure wells and is the cause of large losses of fluid to the forma-
t i ~ n . ~ ~

Backflow: After completion of a treatment, the returns in most cases are flowed back to a pit or tank
as rapidly as possible. The rapid flow carries suspended particles, emulsions, and spent acid that
would be difficult to produce without substantial driving energy. In the case of slow backflow from a
reservoir with low driving energy, the recovery of fluids may be aided by swabbing, nitrogen gas or
CO2 gas in the fluid or preflush, or by artificial lift. Regardless of the recovery method, in most
instances the stimulation fluids should be recovered as rapidly as feasible. Only with some unconsoli-
dated formations or in proven instances of migrating fines caused by high flow rates should a slow
recovery technique be used. Migrating fines may be identified with a laboratory test that plots perme-
ability response with increasing and decreasing driving pressure and flow rate.

Reaction Rate Factors


Temperature
Acid reaction rate on most acid soluble materials increases with temperature because of a lowering of
viscosity. Hydrochloric acid reacts with limestone almost instantaneously and is affected in the tem-
perature range of 60°F (16°C) to 200°F (93°C) only by the transfer of the acid to the formation and the
transfer of soluble by-products away from the reaction site. As the treatment temperatures increase,
the viscosity of the acid is lowered, allowing the acid to move into smaller pores and cracks. Also at
higher temperatures the reaction by-products are more readily soluble in the HCI acid which aids in
the transfer of by-products away from the reaction site.

Pressure
The reaction rate of acid is reduced as the pressure rises from atmospheric to about 500 psi
(3450 kPa). Any pressure increase above 500 psi has very little effect on the rate of reaction of the
acid with the formations. Since nearly all acidizing treatments are performed at pressures over
500 psi, the effects of pressure changes will not enter into the treatment design.

Area Volume Ratio


Area-to-volume ratio is the major factor controlling spending time of the acid on a particular formation.
The area-to-volume ratio is the surface area of the formation which is in contact with a given volume of
acid. The ratio is inversely proportional to the width of the fracture or the pore diameter if the treatment
is a matrix acidizing treatment. Area-to-volume ratios may range up to 20:l in a 1/10in. wide fracture,
200+:1 in an open hairline crack, and on the order of 30,OOO:l in a matrix with a porosity of 20% and a
permeability of 10 md. At high area volume ratios, the acid spending time is short and the penetration
of live acid is greatly reduced. For this reason, live acid will penetrate farther in an open hydrau-
lic fracture than in a hairline fracture or through the matrix.
Acid Concentration
Initially, 15% HCI was picked as a standard acid strength because of the poor performance of the first
inhibitor, sodium arsenate, in higher concentrations of acid. As inhibitors were improved, the higher
concentrations of acids were used in fracturing for their higher dissolving capacity. In matrix treating of
sandstones, 15% HCI is usually the upper concentration limit.

and 10% strength are used as damage removal treatments in sand-


Hydrochloric acids of 5, 7-1/2,
stone formations. These weak acids, when coupled with a surfactant, will frequently remove the acid
soluble damage with much less tubular corrosion than the higher strength acids.

Acid Selection
Selection of the acid for the stimulation is very often dictated by the damage and the composition and
temperature of the formation. In the majority of stimulations, HCI acid is used because of its low price,

12-24
lack of insoluble by-products, and high dissolving capacity. For reaction on clay, silt, and low calcium
content sandstone, a mixture of HCI acid and HF acid is used. In high temperature formations (tem-
peratures above about 220°F) where HCI reacts very quickly and is difficult to inhibit, an organic acid
may be considered as a substitute.

Formation Composition
The composition of the rock is important for determining the method of treatment as well as type and
strength of acid. If a formation has very little carbonate but is rich in clay minerals, then an HCI/HF
acid mixture may be the optimum stimulation for increasing impaired permeability in the immediate
area of the wellbore.

Additives
Acid additives are specially developed chemicals that modify the chemical or physical behavior of the
acid in reactions with produced fluids, the formation, or reaction by-products. The additives may be
surfactants, alcohols, hydrocarbon solvents, salts, polymers, and other compounds. They are formu-
lated to solve a particular problem, yet may have other uses besides the primary function. The brief
additive descriptions that follow are intended to provide a brief look at the individual classes of materi-
als.

Surfactants
Surfactants are multifunction chemicals which are added in small volumes to acid to accomplish a cer-
tain task. Some surfactants may help acid penetrate the formation more easily while others may act as
solubilizing agents or clay stabilizers. The amount and type of surfactant to use depends upon the for-
mation and, in some cases, upon individual well characteristics.

One serious problem with surfactants is adsorption onto clay surfaces in the formation.44 Surfactants
which adsorb heavily should be avoided. This can usually be accomplished by knowledge about the
formation and the particular surfactant and the composition and behavior of backflow fluids after a
treatment.

Surfactants can be classified into four major groups. The division in which a surfactant belongs
depends upon the water-soluble group of the surfactant. The divisions are:

1. Anionic

2. Cationic

3. Nonionic

4. Amphoteric

Anionic surfactants have a negatively charged water-soluble group on the end of the molecule. Exam-
ples are the sulfate and sulfonate compounds. The major applications of anionics are as nonemulsify-
ing, retarding, and cleaning agents.

Cationic surfactants have a positively charged water-soluble group. An example is quaternary ammo-
nium chloride. The major uses of the cationics are as nonemulsifiers, corrosion inhibitors, and bacteri-
cides.

Most nonionic surfactants contain polymers as the water-soluble group and hence have no charge.
Examples of the nonionics are polyethylene oxide and polypropylene oxide. Major uses of nonionics
are as nonemulsifiers and foaming agents.

12-25
Amphoteric surfactants are organic molecules with a water-soluble group which may be either posi-
tive, negative, or have no charge. The particular charge of the water-soluble segment of an amphot-
eric surfactant depends upon the pH of the system. Amphoterics have only very limited usage within
the petroleum industry at the present time.

Mixing surfactants may cause reactions between some surfactants of oppositely charged classes to
the extent of rendering some surfactants inactive or damaging. Mixing of the surfactants cannot be
avoided in most cases, since the corrosion inhibitors are usually cationic, while many other acid addi-
tives are anionic; however, concentration ranges have been established by the service companies for
surfactant mixing so that maximum performance of each surfactant can be achieved.

Fluorocarbon surfactants are usually less adsorptive than normal hydrocarbon base surfactants but
can cause severe emulsion problems if used in the wrong concentrations. Fluorocarbons may work
adequately in the range of 5-100 ppm and are diluted usually with alcohol or water. Over use of any
surfactant may change the behavior of any surfactant and cause damage.

Wettability
Wettability is a term used to indicate whether a formation can be preferentially coated with oil or water.
This information may be determined from laboratory tests on the produced oil. Additives which convert
a formation surface from oil-wet to water-wet will speed considerably the reaction between the acid
and the soluble material. These surface preparations may be cleaners, solubilizers, or other water-sol-
uble materials which can strip oil and water-wet the surface. Reservoirs are characteristically water-
wet or oil-wet (a function of the natural surfactants in the crude oil) and the condition of the reservoir in
question should be known before planning a treatment. Removal of oil film from scale deposits is also
important for speeding the reaction of acid in removing the scales. Surfactant influences how the wet-
ted surface will behave since the water-soluble, charged end of the surfactant is adsorbed on the for-
mation leaving the oil-soluble group to influence wettability. In general, cationics (positively charged)
adsorb on sandstone (negatively charged) and anionics (negatively charged) adsorb on limestones
and dolomites (positively

Clay Control Additives: If a formation contains swelling or disintegrating clays which may cause per-
meability reduction, a clay control process may be warranted. These treatments, generally either poly-
mer,51152zirconium o ~ y c h l o r i d eor
, ~ hydroxy
~ aluminum,54 work in different ways, but all seek to
isolate or stabilize the clays to prevent breakup of the clay platelets. The optimum clay stabilizer and
amount of stabilizer solution will depend upon the characteristics of the individual formations.

Surface Tension Reducers: A surface tension reducer aids in the recovery of fluids by decreasing
the amount of energy necessary to push the fluid past gas and liquid boundaries.

Demulsifying Agents and Nonemulsifying Agents: These chemicals are oil-soluble surface active
agents and are normally carried in a water or acid medium at concentrations of 0.1 % to 5%. Nonemul-
sifying additives prevent the formation of emulsions during an acidizing stimulation and demulsifiers
are designed to break emulsions. Special mixtures of HCI and nonemulsifying agents are marketed as
nonemulsifying acids to stimulate formations containing emulsion-causing crude^.^^?^' Proper concen-
tration of the demulsifier or nonemulsifier in the treatment is critical. Too high a concentration of
demulsifier or nonemulsifier can turn some products into very good emulsifiers. For this reason, circu-
lation of the acid tanks (which remixes the additives) before injection is a must. If a mutual solvent is
used in the treatment, neither a demulsifier or a nonemulsifier may not be needed.

Antisludge Additives: Some crude oils react chemically with hydrochloric acid to form semi-solid
particles referred to as sludge!1s42157 This formation of the acid-oil sludge begins almost immediately
after contact between the crude oil and the acid. The tendency toward formation of sludge is affected
by the concentration of the acid and other variables and is best determined by laboratory testing.
Higher strength acids and low API gravity oils have a greater tendency toward sludge formation.

12-26
Sludge may form in the wellbore or in the formation and can completely plug flow channels in the pro-
ducing formation. Sludge is extremely difficult to remove because it is insoluble in most treating fluids.

Alcohols: By including certain alcohols or certain mixtures of alcohols in an acidizing treatment, it is


often possible to take advantage of their many surfactant qualities at a lower cost than buying a set of
individual surfactants such as a penetrating agent and an emulsion breaker.59i59Also, alcohols do not
adsorb in the formation like most surfactants, hence they remain with the treating solution in the for-
mation. Alcohols are normally used at concentrations of 5% to 40% by volume in the treating fluid. The
advantages of alcohols are quick cleanup with less emulsions and water blocks.

Mutual Solvents
Mutual solvents may be the most useful additive in sandstone acidizing and damage removal treat-
ments. Mutual solvents aid in lowering surface tension, breaking emulsions and water blocks and
have been used to reduce surfactant adsorption in the formation.5s6o The mutual solvent should be
considered whenever the possibility of creating emulsions and water blocks exists.

Acid Thickeners: Thickening of acids may be desirable to control leakoff, to inhibit mixing with other
fluids, or to promote differential etching in f r a c t ~ r i n g . ~The
” ~ ~products are occasionally used in matrix
treating, but are usually reserved for fracturing. Guar-gum polymers and some synthetic polymers are
often used as inexpensive thickeners in acid but since the acid acts as a breaker, their performance is
poor. Gelling agents designed for acid may be polymer, surfactant system or crosslinked polymer. Vis-
cosities of polymer and surfactant gelled acids are about 20 to 50 cp or slightly higher, while the
crosslinked acid viscosity may be over 100 cp.

Inhibitors: Acid inhibitors slow the acid reaction on the steel piping in the well system and usually
keep tubular damage to a m i n i m ~ m . Before
’~ any acid stimulation is planned for well, the condition of
the tubular goods should be known. If the casing and tubing are in poor condition or if the condition is
unknown, the use of a workover string should be considered. All acids used in stimulation should be
inhibited to protect the tubular goods as well as the service trucks and tanks. The inhibition protects
the steel by adhering to the metal and forming a very thin coating which the acid will not readily pene-
trate. This film will break down with time so the acid should be flushed out the lines and well tubular
goods after an acidizing job.

Since corrosion inhibitors are usually cationic surfactants, adsorption in sandstone reservoirs will strip
the acid of inhibitor very rapidly. When live acid is returned from the formation, corrosion on tubulars
can be severe. To minimize this problem, the acid concentration should be only high enough to con-
sume the damage. If live acid will be returned to the wellbore, the acid should be flushed out as rapidly
as possible.

iron Control Agents: Hydrochloric acid reacts with iron in tubing, rust, iron scale, siderite or chlorite,
to form iron compounds such as gelatinous iron hydroxide as the pH of the spent acid rises above 2.0
(a relative acid strength of less than 1% HCI).w$E Sequestering agents are added to acid to prevent
the iron precipitation as the acid spends. To determine if a sequestering agent is needed, samples of
the returned acid from a treatment should be analyzed for total iron content and the source identified.

Suspending Agents: During an acid reaction on any acid-soluble material, fine particles which may
be acid-insoluble or slowly acid-soluble are released and carried through the fracture with the flow of
the acid. When the fluid movement stops or slows sufficiently, the particles may fall out of suspension
and plug the fracture. The use of a suspending agent (normally a soap or foaming surfactant) in the
acid will aid in keeping the particles in solution until produced from the well.

12-27
Wash Design
Wellbore Cleanup and Acid Wash
Removal of some scales, coatings, sludges, and other near-wellbore damage can often be accom-
plished with an acid soak or with low strength acid wash.

The basic procedure for an acid “soak” or “spot” is to run tubing completely below or to a point in the
lower half of the open-hole or perforated completion interval and pump the acid down the tubing,
allowing it to stand over the completion zone. Washing is often done on carbonate formations of high
permeability to reduce cement and drilling mud damage. This is also effective in reducing breakdown
pressure on wells to be fractured by reducing cement and perforation damage. If no packer is used, it
is most important that the fluid density of the acid be the same as the density of the fluid in the hole or
the acid will float up or become dispersed in the wellbore fluid. Acid density can be increased with
weighting materials such as calcium chloride if necessary. After the acid is spotted at the perfs, the
wellbore fluid-acid interface will seek a common level inside the tubing and in the casing tubing annu-
lus if the wellbore fluid is kept fully loaded on the tubing and annulus at the surface.

To “wash” the completion zone, the acid should be pumped past the open hole or perforations by dis-
placing it from the tubing (the acid should not be completely displaced from the tubing; this will mini-
mize mixing with the displacement fluid), allowing the wellbore fluid-acid interface to rise in the
annulus. Then the acid can be allowed to “U-tube” back into the tubing by bleeding back some of the
displacing fluids from the tubing while simultaneously pumping fluid back into the annulus. This dis-
placement of acid back and form from the tubing to the annulus should be done several times, so that
the completion interval is subjected to most of the unspent acid.

After the acid has been washed back and forth several times, the well fluid-acid interface can be dis-
placed to near the top of the completion interval by pumping into the tubing. If a packer is not used,
then the annulus should be shut-in to hold the interface at this level or it can be held at that point by
setting the packer (provided the well does not go on “vacuum”). Then all of the acid below this point
and in the tubing can be pumped into formation with minimum pressures.

Solvent Wash of Injection Wells


In many injection wells there are layers of oil sludge built up over the open-hole or perforations which
prevent the well from taking injection water. These sludge layers often contain too much organic com-
pounds for a mutual solvent and acid to effectively remove. To properly treat these cases, a solvent
such as toluene or xylene may be necessary. Although other solvents can disperse oil, xylene and tol-
uene have been found to be the most consistent at solubilizing or dispersing most of the organic mate-
rials found in injection well sludges.

After washing and acidizing, the wellbore fluids should be backflowed or unloaded with nitrogen, if
possible. There is normally too much debris (sand, silt, undissolved oil and trash) remaining after a
treatment to flush the treatment into the formation (Note: This does not apply in injection wells where
there is no sludge).

In instances of very severe sludge buildup, mechanical scrapers and/or jetting nozzles on tubing have
been used successfully to clean wellbores.

Perforation Breakdown
Perforation breakdown treatments with acid have been used to try to open up the perforations prior to
production or further stimulation. These treatments have had very mixed results. A better approach to
breaking down (opening) perforations is the perforation breakdown tool offered by several companies.
The device is basically a perforated nipple between two packers.35 The packers isolate a few feet of
perforations at a time and an acid is pumped down the tubing and through the perforated nipple to
open the perforations. Although water or oil could be used to open the perfs, the acid can remove any
perforating debris and most cement and mud damage.

12-28
The spacing between the packers should be as short as possible since only a few perforations in each
treated interval will be opened before the fluid pressure is lost. The volume of acid pumped into each
section may vary but 25 gallft is usually adequate. This tool should not be used on perforations within
about 10 ft of any water contact due to the possibility of fracturing the zone into the water. If the treat-
ment has to be shut-in for a time exceeding about six hours, consideration should be given to breaking
down the perfs with KCI or NH4CI water instead of acid.

Extreme Overbalance Perforating


An alternate method of breaking down perforations is to use a very high overbalance of a gas driven
clean fluid at the time of perforating. Details are given in the chapter on perforating.

Matrix Acidizing Design


Before designing an acidizing treatment, review the well’s production history and determine if the
potential benefits are worth the stimulation attempt.36 The success ratio for matrix acidizing treat-
ments in the industry is low - estimates range from 30-50% - and many wells are damaged by poorly
designed or unneeded acid treatments. Heading the list of acidizing failures are wells that are indis-
criminately selected for acidizing because of a wild hope of increasing production. Effective use of
acid treatments requires careful examination of the well and its production history. A commitment to
designing a job on the basis of individual well requirements and controlling the quality of the treatment
will insure better results.

A great many acid jobs fail because the “formation damage” that has been treated is actually faulty
perforations. If the number, size or location of the perforations is in doubt, the well should be reperfo-
rated and tested before being acidized.

Obtaining the Required Information


Figure 12.17 is a worksheet that can be used in designing a treatment and illustrates the type of infor-
mation necessary for treatment design. Section I of the figure must be filled in from well records or
field experience before the acidizing treatment can be designed. The items likely to cause the most
difficulty are average undamaged permeability and permeability of the highest permeability zone. The
average undamaged permeability can be obtained from buildup tests or from an average of core per-
meabilities. A core permeability average will likely be much lower than a permeability calculated from
a buildup test since a core permeability will not reflect the contribution of any natural fractures. If a
core permeability is used, it is best to use results of liquid permeability tests for oil and water wells and
of gas permeability tests for gas wells. A core’s permeability to gas or air is often several fold higher
than its permeability to liquid if there is no correction for slippage and turbulence. A buildup permeabil-
ity probably reflects closely the true permeability of the formation and should be used whenever possi-
ble. Actually, kh is the result from the buildup test and the accuracy of kwill depend on which h is
used; i.e., the perforated height or the total interval thickness. For purposes of this procedure, the net
(perforated) zone height should be used except in severely restricted, limited entry completions where
less than 50% of the net pay has been perforated. In that case, the perforated zone height should be
used. The permeability of the zone of highest permeability is a measure of the permeability of any thief
zones, fractured zones, or leakoff zones that may exist in the formation. If there are no thief zones or
stringers of high permeability, then this value should be set the same as the average undamaged per-
meability (k,,). If there is a history of rapid leakoff from the zone, an approximate permeability may be
calculated from leakoff rates or, if available, from core analyses of the high permeability zone on this
well or immediately offset wells. The well’s stimulation history should be investigated since induced
fractures can act as leakoff zones.

Carbonate content and porosity are available from core data. Zone thickness and formation tempera-
ture are available from log data. The type of damage in the rock is best ascertained by examining well
and workover histories. If there is no damage but the formation does respond readily to acid, the for-
mation can still be matrix acidized; however, matrix acidizing an undamaged formation will increase
the production rate only slightly.

12-29
I. Information Needed
formation temperature F
avg. undamaged permeability md
perm. of high perm zone md
carbonate content %
type of damage
avg. porosity (net pay) decimal
net pay thickness ft
tbg. vol + csg below packer bbl
open hole diam. if not cased in.

II. Design Information

1. type ofdamage (Figure 12.19)

2. type of acid (Figure 12.18)


if damage not acid soluble,
select a solvent

3. type of solvent (Figure 12.19)

4. volume of acid/solvent
acid vol. (Figure 12s.18)
(Figure 12.20)
solvent vol. (Figure 12.20)

5. additives (Figure 12.21)


additives (Figure 12.22)
additives (Figure 12.23)
additives

6. stages and diverting


a.no. stages (Figure 12.24)
b.type diverter (Figure 12.25)
c.divt. quant. (Figure 12.26)
d.Vol. stages (Figure 12.26)
e.S.1. P./balls (Figure 12.27)
f S. I.P./spacing (Figure 12.27)

7. Nitrogen or CO gas use?


decision based on (Figure 12.28)
volumes set after contact with
service company engineer

8. Ovemush
ovemush type (see page 34)
vol. of fluid (Figure 12.29)

Figure 12.17: Chemical Stimulation Design Work Sheet

12-30
Designing the Treatment
Section I I of the worksheet is filled out with the help of Figures 12.18 through 12.27.

The acid is usually selected (Figure 12.18) on the basis of the formation mineralogy, the formation
temperature and the type of damage in the sandstone. For sandstones with total carbonate content
below 10-15%, the best acid at temperatures below 250°F is usually HCVHF (HCVHF acid in this
report refers exclusively to 12% HCI + 3% HF). At formation carbonate concentrations greater than
15%, acids containing HF are usually not recommended because an insoluble precipitate - calcium
fluoride, CaF2 - may form. At temperatures greater than 250"F, the HCI is typically replaced with for-
mic acid or acetic acid.

~~

Less tll ian 5% 5% to 10% 10% to 15% I Over 15%


Temperature
(" F)
Treatment Temperature
m I Treatment I I Treatment

QOO 50 ga'/n
15% HCI
followed by by perforation
HCllHF wash with
HCVHF
200-250 35 gallit 200-250 50 gaVft 15%
10% HCI HCI followed
followed by by HCllHF
HCVHF
250-350 35 gaVft 250-350 35 gaVft 258350 50 gaVft 7.5% 250-350 50 gal/ft
10% formic 7.5% HCI & HCI & 10% 10% 7.5%
acid, for- 10% formic formic acid HCI & 10%
miclHC1, or acid fol- formic acid
formic HF lowed by for-
rnidHF
>350 35 galin >350 35 gallft 10%
10% formic formic acid formic acid 10% formic
acid, or for- followed by acid
mic/HF formicRlF

Figure 12.18: Carbonate Content

Figure 12.19 lists common types of formation damage and recommended solvents for treating them.
The type of damage may be what ultimately dictates which acid is used. However in dealing with insol-
uble precipitates, other limitations may take precedence (for example, even if there is drilling mud
damage, HCI/HF should not be used if the sandstone is limy; a better approach is a small fracture
treatment). The volume of acid (Figure 12.20) used in a matrix acid treatment varies with the perme-
ability. (The depth of damage is also important; however it usually cannot be determined.) In low per-
meability zones where injection is very slow, it is advisable to use small volumes of acid to avoid
(1) corroding the tubing because of long acid residence time, (2) fracturing the sandstone with the
acid, or (3) precipitating acid reaction products out of the spent acid during the long period necessary
for recovering the spent acid. On formations with permeabilities less than 0.1 md, acid is recom-
mended only for perforation breakdown. Acid may even be replaced here by a clean 2% KCI water

Selecting additives is the biggest challenge in treatment design. Figures 12.21 through 12.25 list rec-
ommended additives and surfactants for various conditions. Selecting some additives such as mutual
solvents precludes using other additives such as demulsifiers, nonemulsifiers, alcohols, or surface
tension lowering surfactants. The use of other additives, such as iron sequestering additives depends
upon the amount of rust or iron scale in the tubing. If excessive rust or iron scale is present, a pretreat-
ment cleanout may be required. Do not use clay control additives unless the formation has demon-
strated a sensitivity to produced waters or to waters that will be used in the treatment. Clay control

12-31
additives often reduce permeability by as much as 50%. There is a common test that purports to show
water sensitivity by injecting a one normal sodium chloride brine and following it with distilled water.
However unless sodium chloride brine followed by fresh water is to be used in the well, it is not a valid
test to demonstrate need for a clay control treatment. Sometimes in the area near the wellbore in an
unfractured well, the increase in velocity caused by converging radial flow will result in the production
of formation fines that can reduce the permeability. In such cases, a clay-control additive may be use-
ful. If the fines are moving throughout the entire formation, however, a clay control treatment will not
be effective.

Figure 12.19: Solvents for Common Damage Conditions in Sandstones

Nitrogen gas may be useful if the formation is a gas zone with permeability below about 20 md and is
not naturally fractured. Whether to use nitrogen gas in the treatment can be decided with the aid of
Figure 12.26. The nitrogen supply company can help decide how much nitrogen gas to use. The vol-
ume will be based on bottomhole pressure, depth and the size of the treatment.

12-32
Figure 12.20: Suggested Acid Volumes

Additive Use
Corrosion Inhibitor always used when acidizing
Clay Control Agent use only if problem clays may exist
Friction Reducer use in high rate fracturing treatments
' Silt Suspender when clearing drilling mud or when acid reaction produces more than 10% silt
Iron Control Agent Use only if analysis of previous acid backflow shows >1500 ppm total iron. Also
needed where >10% chlorite, magnetite or siderite exists. Rusty tbg. must be
cleaned or redaced before acid.

Figure 12.21: Specific Purpose Acid Additives

The overflush volume is usually minimized since acid spends fairly rapidly in the formation matrix and
a large overflush would unnecessarily contribute more load fluid to recover. The recommended over-
flush volumes are given in Figure 12.27. The type of overflush fluids depends upon the acid and the
formation. For HCI treatments, one can use filtered produced water, 2% KCI water, 2% NH4CI water,
gas (in a gas well), or oil (in an oil well) - as long as the fluid is CLEAN. For HCllHF acid, acceptable
overflushes are clean 2% NH4CI water, oil, gas or HCI. If an HCI afterflush is used, 5% to 7-112% HCI
is usually adequate.

12-33
I Additive I Use I
1
I

Miscible/mutual Solvent I use when emulsions and sludges- .present may- replace
. many surtactants
Alcohol to remove water blocks, helps recover water or spent acid, breaks some
emulsions
Antisludge prevents sludge in oils where a proven sludge tendency exists
Nonemulsifier prevents emulsions, testing mandatory not needed when mutual solvent is
used
Demulsifier breaks emulsions, testing mandatory not needed when mutual solvent is
used

Figure 12.22: General Purpose Acid Additives

Additive Use
Mutual solvent neips remove water blocks in oil zones
Surf. Tens. lowering helps remove water blocks in oil zones
Dispersants helps solvent penetrate paraffin and sludges
Foamers used in combination with gas to help unload well

Figure 12.23: Solvent Additives

Average Undamaged
Permeability X Zone Recommendations
Thickness*, k,h
10 to 100 md tt Two stages; put diverter in last third or tirst stage, or use perforation
wash tool.
100 to 1000 md ft Three stages; put diverter in last third of stages one and two, or use
perforation wash tool.
>loo0 md ft or natural fractures Four stages; put diverter in last third of stages one, two, and three
or use perforation wash tool.

Figure 12.24: Number of Stages for a Matrix Acid Treatment

After the worksheet has been completed, pressure calculations will need to be added and a workover
form completed.

12-34
Thermal stimulation may be accomplished by electric resistance heating, hot oil, water injection or
controlled circulation, gas burners, in-situ combustion, exothermic reactions or steam injection.

Electric heat generation devices have been used in areas producing viscous crudes to lower the oil
viscosity or where the paraffin cloud point (precipitation point) is nearly the same as the bottom hole
temperature. Both continuous and intermittent heaters are available for specific applications. The total
quantity of wells heated by electrical resistance heating is small; the expense of electricity and power
losses are the chief reasons.

The response from wells in California and the USSR shows varied response depending upon sand
thickness and amount of damage in the reservoir. In California, a 13" API crude oil with a viscosity of
over 3500 centipoise at the wellbore temperature of 80°F was successfully produced after the temper-
ature at the wellbore was increased to 140" (viscosity dropped to 210 centipoise). Payout time of the
capitol cost of equipment was approximately two years.

Electric heaters currently available may be of several types. Usually the power ratings are between 9-
60 kilowatts (20,000-500,000 BTU/hour). These heaters may deliver heat generation in excess of
2OO0F, although the depth penetration of the heat into the wellbore is limited, especially as fluids flow
toward the wellbore.

Gas burners are suitable for heating wellbores to much higher temperatures and for initiating fire
floods and fire flood clean up around the wellbore. A gas burner involves injection of gas down the
tubing while air is injected down the casing/tubing annulus. The gas is ignited by an ignition system at
the tool creating extremely high temperatures. If the zone is a fire flood, the flame is exposed to the
formation and if the thermal properties of the gas are to be exploited without a fire flood, the tool uses
a shield to protect the casing from heat damage. Heat is transferred by air flowing around the heat
shield and into the formation.

Hot water circulation is a simple method of wellbore heating, but it is not usually effective because the
tubing and casing become a shell and tube exchanger: the returning fluid robs heat from the injected
fluid. Methods that make the process work are insulated tubing and dual tubing strings. Bull heading
(no circulation) is also an effective method of getting heat to bottom hole although formation damage
is severe.66

Stimulation of oil wells by combustion is an outgrowth of fire flood technology. The familiar combustion
front is initiated at the wellbore by a burner and propagated into the formation to a distance of approx-
imately 10-20 ft. The cleaned formation surrounding the wellbore serves as a pathway of improved
permeability to the returning fluids when the well is put back on production.

Hot water and steam injection provide effective well stimulation. Steam is the more widely used of any
of the high capacity thermal methods and is capable of millions of BTU input into a formation. Use of
hot water in stimulations have met with only limited success because of heat transfer problems. Cyclic
steam injection over a period of weeks or months may provide roughly the same productivity increase
as a fire flood.

Chemical Heat Generation


Heat may also be produced either downhole or at the surface by the use of exothermic reaction^.^^-^'
The most common heat producer involves the reaction of sodium nitrite with another salt, such as
ammonium nitrate or ammonium chloride, to produce heat and nitrogen gas?8 The advantage of this
reaction over other exothermic reactions is that it may be controlled by buffers that affect the pH of the
system. By control of the mixing and the buffers, the point at which maximum heat is generated can be
predicted and controlled. Maximum temperatures recorded with the process are in excess of 400°F
and the total heat available depends upon the volume of the reactants. The process offers some treat-
ing versatility, since the buffered reactants can be combined with aromatic solvents in a slightly stabi-

12-37
lized emulsion that will provide a heated solvent at a predetermined place in the ~ e l l b o r e . ~
The
’ main
application in the technical literature has been in paraffin removal.

A second heat generation process, specifically for acid, is addition of ammonia to hydrochloric acid.
The reaction of the acid with the ammonia generates an immediate heat rise.71The need for heating
hydrochloric acid is rare but has application in fracturing a cool, low reactivity dolomite, prevention of
paraffin precipitation, removing high viscosity oil coatings from an acid reactive surface, as an aid in
breaking emulsions, and to minimize tubing contraction caused by injecting cool acid in a hot well.

References
1. Williams, B. B.: “Fluid Loss from Hydraulically Induced Fractures,” Journal of Petroleum Technol-
ogy, (July 1970), 882-888.

2. Harrison, N. W.: “Diverting Agents - History and Application,” Journal of Petroleum Technology
(May 1972), 593-598.

3. Ellenberger, C. W. and Aseltine, R. J.: “Selective Acid Stimulation to Improve Vertical Efficiency
in Injection Wells - A Case History,” Journal of Petroleum Technology (Jan. 1977), 25-29.

4. Best, B. W. and Miller, L. 0.: “Optimum Use of Diverting Agents in Well Stimulation Treatments,”
Stimulation, 101-103.

5. Webster, K. R.; Goins, W. C., Jr. and Berry, S. C.: “A Continuous Multi-stage Fracturing Tech-
nique,” Journal of Petroleum Technology (June 1965), 619-625.

6. Stipp, L. C. and Williford, I?. A.: “Pseudolimited Entry: A Send Fracturing Technique for Simulta-
neous Treatment of Multiple Pays,” Journal of Petroleum Technology (May 1968), 457-462.

7. Lagrone, K. W. and Rasmussen, J. W.: “A New Development in Completion Methods-The Lim-


ited Entry Technique,” Journal of Petroleum Technology (July 1963), 695-702.

8. Streltsova-Adams, T. D.: “Pressure Drawdown in a Well with Limited Flow Entry,” Journal of
Petroleum Technology (Nov. 1979), 1469-1476.

9. Jones, L. G. and Slusser, M. L.: “The Estimation of Productivity Loss Caused by Perforation -
Including Partial Completion and Limited Entry,” Paper SPE 4798, presented at the Second Mid-
west Oil and Gas Symposium, Indianapolis, March 28-29, 1974.

10. Hushbeck, D. F.: “Precision Perforation Breakdown for More Effective Stimulation Jobs,” Paper
SPE 14096, Int. Mtg. Pet. Eng. Beijing, March 17-20, 1986.

11. Ande, T. J. and Perkins, D. B.: “Thru-Tubing Plugback Tools and Services,” Amoco New Orleans
Engineering Report, November 16, 1987.

12. Brown, R. W.; Neill, G. H. and Loper, R. G.: “Factors Influencing Optimum Ball Sealer Perfor-
mance,” Journal of Petroleum Technology, (April 1963), 450-454.

13. Crurnp, J. B. and Conway, M. W.: “Effects of Perforation-Entry Friction on Bottom Hole Treating
Analysis,” Journal of Petroleum Technology (Aug. 1988), 1041-1048.

14. Erbstoesser, S. R.: “Improved Ball Sealer Diversion,” Journal of Petroleum Technology, (Nov.
1980), 1903-1910.

12-38
15. Gabriel, G. A. and Erbstoesser, S.R.: “The Design of Buoyant Ball Sealer Treatments,” Paper
SPE 13085 presented at the 59th Annual Technical Conference and Exhibition, Houston,
September 16-19, 1984.

16. King, G. E. and Hollingsworth, F. H.: “Evaluation of Diverting Agent Effectiveness and Cleanup
Characteristics Using a Dynamic Laboratory Model - High Permeability Case,” Paper SPE 8400,
54th Annual Fall Mtg, Las Vegas, September 23-26, 1979.

17. Hill, A. D. and Galloway, P. J.: “Laboratory and Theoretical Modeling of Diverting Agent Behav-
ior,” Journal of Petroleum Technology (July 1984), 1157-1163.

18. Houchin, L. R., Dunlap, D. D.,Hudson, L. M. and Begnaud, P. C.: “Evaluation of Oil-Soluble
Resin as an Acid-Diverting Agent,” Paper SPE 15574 presented at the 61st Annual Technical
Conference and Exhibition of the Society of Petroleum Engineers, New Orleans, October 5-8,
1986.

19. Coulter, A. W., Crowe, C. W., Barrett, N. D. and Miller, B. D.: “Alternate Stages of Pad Fluid and
Acid Provide Improved Leakoff Control for Fracture Acidizing,” Paper SPE 6124 presented at the
51st Annual Fall Technical Conference and Exhibition of the Society of Petroleum Engineers of
AIME, New Orleans, October 3-6, 1976.

20. Dill, W. R.: “A Gel Diverting Agent Used in Acidizing Treatments,” Halliburton Services, Produc-
tion Engineering, 1978, 111-115.

21. Burman, J. W. and Hall, B. E.: “Foam as a Diverting Technique for Matrix Sandstone Stimula-
tion,” Paper SPE 15575 presented at the 61st Annual Technical Conference and Exhibition of the
Society of Petroleum Engineers, New Orleans, October 5-8, 1986.

22. Penny, G. S.: “Nondamaging Fluid Loss Additives for Use in Hydraulic Fracturing of Gas Wells,”
Paper SPE 10659 presented at the SPE Formation Damage Control Symposium, Lafayette,
March 24-25, 1982.

23. King, G. E.: “Foam and Nitrified Fluid Treatments - Stimulation Techniques and More,” Paper
SPE 14477 presented as a Distinguished Lecture during the 1985-86 SPE Distinguished Lec-
turer Program.

24. Schriefer, F. E. and Shaw, M. S.: “Use of Fine Salt as a Fluid Loss Material in Acid Fracturing
Stimulation Treatments,” Paper SPE 7570 presented at the 53rd Annual Fall Technical Confer-
ence and Exhibition, Houston, October 1-3, 1978.

25. Miller, B. D. and Warembourg, P. A.: “Prepack Technique Using Fine Sand Improves Results of
Fracturing and Fracture Acidizing Treatments,” Paper SPE 5643 presented at the 50th Annual
Fall Meeting, Dallas, September 28-October 1, 1975.

26. Fry, W. C. D., Boney, C. L., Atchley, J. W. and Whitsett, F. T.: “The Use of 100 Mesh Sand for
Improving Acid Efficiency,” Paper presented at the Southwest Petroleum Short Course, Lubbock,
79-82.

27. Dill, W. R.: “Effect of Bridging Agents and Carrier Fluids on Diverting Efficiency,” Journal of
Petroleum Technology (Oct. 1969), 1347-1352.

28. Pye, D. S.and Smith, W. A.: “Fluid Loss Additive Seriously Reduces Fracture Proppant Conduc-
tivity and Formation Permeability,” Paper SPE 4680 presented at the 48th Annual Fall Meeting of
the Society of Petroleum Engineers of AIME, Las Vegas, September 30 - October 3, 1973.

12-39
29 Canson, B. E.: “Lost Circulation Treatments for Naturally Fractured, Vugular, or Cavernous For-
mations,” Paper SPE/IADC 13440 presented at the SPE/IADC 1985 Drilling Conference, New
Orleans, March 6-8, 1985.

30. Coberly, C. J. and Wagoner, E. M.,: “Some Considerations in the Selection and Installation of
Gravel Packs for Oil Wells,” Journal of Petroleum Technology (Aug. 1938), 1-20.

31. Mahajon, N. C. and Barrow, B. M.,: “Bridging Particle Size Distribution: A Key Factor in the
Design of Non-Damaging Completion Fluids,” SPE 8792, 4th Symposium on Formation Damage
Control, Bakersfield, January 28 & 29, 1980.

32. Super X Acid Technical Report, Dowell, DWL1313-26M-966.

33. McBride, J. R., Rathbone, M. J., and Thomas, R. L.: “Evaluation of Fluoroboric Acid Treatment in
the Grand Isle Offshore Area Using Multiple Rate Flow Test,” Paper SPE 8399 presented at the
54th Annual Fall Technical Conference and Exhibition, Las Vegas, September 23-26, 1979.

34. Gdanski, R.: “AICI3 Retards HF Acid for More Effective Stimulations,” Oil and Gas Journal (Octo-
ber 1985), pp. 111-116.

35. Jefferies-Harris, M. J., Coppel, C. P.: “Solvent Stimulation in Low Gravity Oil Reservoirs,” JPT
(February 1969), pp. 165-175.

36. King, G. E., Holman, G. B.: “Hydrocarbon Solvents: An Alternative to Acid for Removing Some
Formation Damage,” SPE 14136, presented at the 1986 Internat. Mtg. on Pet. Eng., Beijing,
March 17-20, 1986.

37 * Douglass, B. C., King, G. E.: “A Comparison of Solvent/Acid Workovers in Embar Completions -


Little Buffalo Basin Field,” SPE 15167, presented at the Rocky Mountain Regional Mtg., Billings,
MT, May 19-21, 1986.

38. Minter, R. B., Davis, E. E., Conway, E. E.: “An Acid-Solvent Stimulation Technique for Low Grav-
ity Crudes,” SPE 3189, 41st Calif. Reg. Mtg., October 28-30, 1970.

39. Gidley, J. L.: “Acidizing Sandstone Formations - A Detailed Examination of Recent Experience,”
Paper SPE 14164 presented at the 60th Annual Technical Conference, Las Vegas, September
22-25, 1985.

40. King, G. E., Holman, G. B.: “Quality Control at Well Site Optimizes Acidizing Economics,” Oil and
Gas J. (March 18, 1985), pp. 139-142.

41. Jacobs, I.: “Asphaltene Precipitation During Acid Stimulation Treatments,” SPE 14823, Forma-
tion Damage Symposium, Lafayette, 1986.

42. Moore, E. W., Crowe, C. W., Henrickson, A. R.: “Formation Effect and Prevention of Asphaltene
Sludges During Stimulation Treatment,” J. Pet. Tech. (September 1965), pp. 1023-1028. J

43. Holditch, S. A.: “Factors Affecting Water Blocking and Gas Flow from Hydraulically Fractured
Gas Wells,” Paper SPE 7561 presented at the 53rd Annual Fall Technical Conference and Exhi-
bition of the Society of Petroleum Engineers of AIME, Houston, October 1-3, 1978.

44. King, G. E., Lee, R. M.: “Adsorption and Chlorination of Mutual Solvents Used in Acidizing,” SPE
Prod. Eng. (May 1988), pp. 205-209.

12-40
45. Anderson, W. G.: “Wettability Literature Survey - Part 1: Wettability Measurement,” Journal of
Petroleum Technology (November 1986), pp. 1246-1262.

46. Anderson, W. G.: “Wettability Literature Survey - Part 2: Wettability Measurement,” Journal of
Petroleum Technology (November 1986), pp. 1246-1262.

47. Anderson, W. G.: “Wettability Literature Survey - Part 3: Wettability Measurement,” Journal of
Petroleum Technology (November 1986), pp. 1246-1262.

48. Anderson, W. G.: “Wettability Literature Survey - Part 4: Wettability Measurement,” Journal of
Petroleum Technology (November 1986), pp. 1246-1262.

49. Anderson, W. G.: “Wettability Literature Survey - Part 5: Wettability Measurement,” Journal of
Petroleum Technology (November 1986), pp. 1246-1262.

50. Anderson, W. G.: “Wettability Literature Survey - Part 6: Wettability Measurement,” Journal of
Petroleum Technology (November 1986), pp. 1246-1262.

51. Woodroof, R. A. and Anderson, R. W.: “Synthetic Polymer Friction Reducers Can Cause Forma-
tion Damage,” Paper SPE 6812 presented at the 52nd Annual Fall Technical Conference and
Exhibition of the Society of Petroleum Engineers of AIME, Denver, October 9-12, 1977.

52. McLaughlin, H. C., Elphingstone, E. A., and Hall, B. E.: “Aqueous Polymers for Treating Clay in
Oil and Gas Producing Formations,” SPE 6008, New Orleans, October 3-6, 1976.

53. Veley, C. D.: “How Hydrolyzable Metal Ions React with Clays to Control Formation Water Sensi-
tivity,” J. Pet. Tech. (September 1969), pp. 1111-1118.

54. Reed, M. G.: “Stabilization of Formation Clays with Hydroxy Aluminum Solutions,” JPT, July
1972.

55. Gidley, J. L. and Hanson, H. R.: “Prevention of Central Terminal Upsets Related to Stimulation
and Consolidation Treatments,” SPE 4551, 48th Annual Fall Mtg., Las Vegas, September 30-
October 3, 1973.

56. Coppel, C. P.: “Factors Causing Emulsion Upsets in Surface Facilities Following Acid Stimula-
tion,” JPT, September 1975, pp. 1060-1066.

57. Delorey, J. R. and Taylor, R. S.: “Recent Studies Into Iron/Surfactant/Sludge Interactions in
Acidizing,” Paper Petroleum Society of CIM 85-36-38 presented at the 36th Annual Technical
Meeting of the Petroleum Society of CIM Held Jointly with the Canadian Society of Petroleum
Geologists, Edmonton, June 2-5, 1985.

58. Gidley, J. L.: “Stimulation of Sandstone Formations with the Acid-Mutual Solvent Method,” J. Pet.
Tech. (May 1971), pp. 551-558.

59. Hall, B. E.: “The Effect of Mutual Solvents on Adsorption in Sandstone Acidizing,” JPT (Decem-
ber 1975), pp. 1439-1442.

60. King, G. E., Brown, T. M.: “Performance of Amoco A-Sol as a Mutual Solvent System,” SWPSC,
April 1978, Lubbock.

12-41
61. Crowe, C. W., Martin, R. C., and Michaelis, A. M.: “Evaluation of Acid Gelling Agents for Use in
Well Stimulation,” Paper SPE 9384 presented at the 55th Annual Fall Technical Conference and
Exhibition, Dallas, September 21-24, 1980.

62. Pabley, A. S. and Holcomb, D. L.: “A New Stimulation Technique: High Strength Crosslinked
Acid,” Paper SPE 9241 presented at the 55th Annual Fall Technical Conference and Exhibition,
Dallas, September 21-24, 1980.

63. Church, D. C., Quisenberry, J. L., and Fox, K. B.: “Field Evaluation of Gelled Acid for Carbonate
Formations,” Journal of Petroleum Technology (December 1981), pp. 2471-2473.

64. Gougler, P. D., Hendrick, J. E., and Coulter, A. W.: “Field Investigation Identifies Source and
Magnitude of Iron Problems,” Paper SPE 13812 presented at the SPE 1985 Production Opera-
tions Symposium, Oklahoma City, March 10-12, 1985.

65. Smolarchuk, P. and Dill, W.: ‘;Iron Control in Fracturing and Acidizing Operations,” Paper Petro-
leum Society of CIM 86-37-28 presented at the 37th Annual Technical Meeting of the Petroleum
Society of CIM, Calgary, June 8-11, 1986.

66. Antheunis, D, Davies, D. R and Richardson, E. A.: “Field Application of In-Situ Nitrogen Gas
Generation System,” Paper SPE 9653 presented at the Middle East Oil Technical Conference of
SPE; Manama, Bahrain; March 9-12, 1981.

67. Collesi, J. B., Donavan, S.C., McSpadden, H. W. and Mitchell, T. I.: “Field Application of a
Chemical Heat and Nitrogen Generating System,” Paper SPE 12776 presented at the 1984 Cal-
ifornia Regional Meeting, Long Beach, April l 1-13, 1984.

68. McSpadden, H. W., Tyler, M. L. and Velasco, T. T.: “In-Situ Heat and Paraffin Inhibitor Combina-
tion Prove Cost Effective in NPR #3, Casper, Wyoming,” Paper SPE 15098 presented at the 56th
California Regional Meeting of SPE, Oakland, April 2-4, 1986.

69. Ashton, J. P., Credeur, D. J., Kirspel, L. J. and Nguyen, H. T.: “In-Situ Heat System Stimulates
Paraffinic Crude Producers in Gulf of Mexico,” Paper SPE 15660 presented at the 61st Annual
Technical Conference and Exhibition of SPE, New Orleans, October 5-8, 1986.

70. Collesi, J. B., McSpadden, H. W. and Scott, T. A.: “Surface Equipment Cleanup Utilizing In-Situ
Heat,” Paper SPE 16215 presented at the SPE Production Operations Symposium, Oklahoma
City, March 8-10, 1987.

71. Hoch, O., Fredrickson, S., Norman, L. and Walker, M. L.: “Heated Acids for Improved Stimulation
Results,” Paper CIM 86-37-68 presented at the 37th Annual Technical Meeting of the Petroleum
Society of CIM, Calgary, June 8-1 1, 1986.

72. Straub, T. J., Autry, S. W., King, G. E.: “An Investigation Into Practical Removal of Downhole Par-
affin by Thermal Methods and Chemical Solvents,” SPE 18889, Production Operations Sympo-
sium, Oklahoma City, March 13-14, 1989, pp. 577-584.

12-42
-
Appendix 12.A Diverters and Fluid Loss
Granular Salt
The 100-mesh salt product is a range of salt sizes varying between approximately 70 and 140-mesh.
Salt is one of the easiest carried diverters and can be used in any well which produces water or in any
well which can be flushed with water.24 The salt may be pumped only in oil, saturated brines, in con-
centrated acids, and in weak acids which have been salt saturated. A graph of the solubility of salt in
acid is shown in Figure 12.30. Once placed, the salt is soluble in almost all produced waters, most
spent acids and aqueous overflushes. The 100-mesh salt is most effective on perforations and natural
fracture sand is easily removed, Figure 12.31.

Average Percent of Initial Permeabiiity


Diverter Regained
After First 1/4" of Injection Face Removed
100 Mesn Sand 100%
Granulated Salt (no polymer) 100%
Resin Dispersion '92%
Benzoic Acid 82%
Naphthalene 77%
Polymer 74% (Typical)

Figure 12.31: Penetration and Removal of Various Diverters

100-Mesh Sand
The 100-mesh sand, sometimes referred to as Oklahoma No. 1, has a size distribution roughly
between 70- and 140-mesh. This sand, which has been used in fracturing treatments, is not an effec-
tive propping agent but rather a fluid loss or diverter materia1.25~26The sand is carried by acid, brine,
oil, water, or gelled water, and is a good diverting agent for perforations or natural fractures. Sand is
not recommended in matrix acidizing of an unfractured well since it cannot be removed from the perfo-
ration by any assurable method. The use of 100 mesh sand is generally discouraged due to damage
on pumps as it is produced back. Where it is required, a temporary pump is usually run for one to two
weeks after the treatment to allow the well to clean up.

Calcium Carbonate
Calcium carbonate is available in sizes from crushed oyster shells and pea size granules to powder.
This material is a fill type diverter that can be used in high rate chemical matrix treating and then later
removed with acid. It can be placed with a gelled brine. It is not recommended for diverting in matrix
acidizing treatments since it is very rapidly soluble in HCI. Calcium carbonate comes in a wide range
of sizes, Figure 12.32. The most effective size range to bridge and seal is from 1/6 the pore size to
about 7 times the pore size.

Naphthalene
Naphthalene flakes are a whitish, thin flake-type diverter with a density of 1.2. They are soluble in
xylene, toluene, condensate, and in a gas stream by sublimation. Naphthalene has been used suc-
cessfully in reservoirs to divert acid; however, flake type diverters should not be used in loadings over
1 Ib/perforations or where there are very small perforations (such as through tubing perforations). In
this situation, the flake type diverters may jam together and be extremely difficult to unload when the

12.A-43
well is turned around for backflow. Naphthalene flakes may be placed by acid or water but should not
be placed with a polymer water. The polymer coats the naphthalene flake and makes it very difficult to
dissolve the flake in its normal solvents. If naphthalene flakes are used, a small overflush (15-
20 gal/ft) of xylene, toluene, or condensate is recommended to help remove the material. Without the
solvent, naphthalene is still removable by gas, Figure 12.33, although only the material in the path of
the gas can be removed. Removal by sublimation is also dependent upon the temperature of the gas
stream.

Benzoic Acid Flakes


Benzoic acid flakes, which physically resemble naphthalene flakes, are also limited in their use to
wells with relatively large perforations or natural fractures. Benzoic acid flakes are soluble in water,
acid, oil, and in gas by sublimation. They may be placed by acid or water but not by polymer water for
the same reason as naphthalene flakes. Benzoic acid flakes should not be used in gas wells with a
temperature below 120°F since the removal by gas sublimation at low temperature is extremely slow,
Figure 12.33.

Wax Beads
Wax beads are small wax pellets. They may be placed by water or acid solutions and can only be
removed by hydrocarbons at certain temperatures. Solubility in oil follows the softening step at tem-
perature. When using this material, select an oil softening point at least 20°F below the static bottom-
hole temperature. Although these materials can be easily transported by water or acid and are
effective in fluid control, Figure 12.34, their density of 0.8 causes them to float and makes them inef-
fective in most matrix operations where pump rate is not in turbulence.

Organic Resin Beads


Organic resin beads have a size range of -70 to +140 mesh. They are available as a dry additive that
may be placed with acid and are effective in slowing the acid rate into a perforation or a fracture. The
beads are soluble in xylene, toluene; condensate, and very slowly soluble in oil. They may be placed
in any aqueous solution and are normally run at the rate of about 1/4 to 1/2 Ib per perforation, or
approximately 1 to 2 Ib/ft of open hole. The density is 1.04, thus they can be easily transported in
water or matrix treating acids and can be used in either normal matrix treating, interface treating or in
any non-oil fracturing fluid. The organic resin diverters cannot be used where a mutual solvent or aro-
matic-acid dispersion is used since the diverter would be prematurely dissolved. The resin beads can
be used in mixtures of methyl or isoprdpyl alcohol and acid or water. A small hydrocarbon overflush
may be useful after an acid job where beads are used.

Organic Resin Dispersions


The organic resin dispersions are effective as a fluid loss agent in oil or wet gas wells where there are
no significant natural fractures. If there are natural fractures, very large quantities of this material
would be required for fluid loss control since it does not bridge the natural fractures. The material can
be placed by either water or acid solution and is soluble in any hydrocarbon production or overflush. A
small hydrocarbon overflush (1 5-20 gal/ft) of xylene, toluene, or condensate can be used after the
acid treatment to speed cleanup.

Micron Size Particulates


Very rapid fluid loss control may be achieved by the micron size particulate diverters. The particles are
small enough to be carried with the placement fluid (acid, water or oil) as a dispersion and can be
used for any matrix operations in formations which are not significantly naturally fractured. These
materials are effective in reducing fluid loss, however, they do create a significant amount of formation
damage since they do not have a solvent.

12.A-44
Polymers
Non-crosslinked polymers control leak-off by viscosity control and are often associated with other
types of fluid loss control. These materials, which are usually guar or a chemically modified guar, are
effective in controlling fluid loss in formations, Figure 12.35, but may achieve fluid loss control by
building a filter cake of polymer debris (wall building). At higher permeabilities, crosslinked polymer
gels should be considered. Polymer usage generally ranges in concentrations from 10 to approxi-
mately 50 Ibs/lOOO gal. Acid solutions can be gelled by special polymers, but these materials are usu-
ally not used in matrix acidizing since they reduce the acid flow into all zones.

Others
Materials such as paper, shredded cloth or leather, grain, or sawdust are not recommended since per-
manent permeability damage is produced.

12.A-45

Você também pode gostar