Вы находитесь на странице: 1из 15

Fluid Functions

The functions of fracturing fluids are to

initiate the fracture

propagate or extend the fracture

transport the proppant where required

return to the wellbore after the treatment

These functions call for unique, extreme and sometimes contradictory fluid
specifications. Depending on what is required from a fluid at various treatment
stages, it may be advantageous for a given material property to exhibit varying
values.

Fracture Initiation

The primary function of the first fluid pumped is to initiate the fracture, by increasing
the pore pressure in the rock until the tensional stress limit is exceeded and rupture
occurs. Its primary performance criterion, therefore, is a high leakoff rate. This allows
it to enter the pores and increase the pore pressure, and also to overcome the high
stress concentrations present around the wellbore due to damage caused during
drilling. To satisfy this requirement, the fluid must be what is termed a "penetrating
fluid" (acid with no fluid-loss control is an example of an excellent penetrating fluid).
Using a non-penetrating fluid to initiate a fracture would cause a higher breakdown
pressure. Weak acids and methanol-brine mixtures are typically used as breakdown
(pre-pad) fluids.

Fracture Propagation

The second function of the fracturing fluid is to extend or propagate the fracture after
it has been initiated. The less the fluid leaks off, the larger the fracture will be. Fluid
leakoff control is therefore the key to fracture propagation. The next function of the
fluid is to ensure that the created fracture is wide enough to accept the injected solid
propping agent, and to transport this proppant to the desired location. Since viscous
energy dissipation and created fracture width are interconnected, the fluid viscosity
(its rheological behavior) is of primary importance for creating fracture width.

Proppant Transport

Propping agents characteristically have a higher specific gravity than the carrier fluid,
and therefore tend to settle to the bottom of the fracture. Sufficiently high fluid
viscosities will enhance the fluid’s proppant-carrying capability. In considering
proppant transport, we should be aware of several factors affecting viscosity. In
particular, the temperature change as the fluid spends more and more time in the
fracture and the particular shear rate at a given location in the fracture both affect
viscosity and hence, proppant-carrying capacity.

Cleanup
Thus far, we have demonstrated that

the first fluid pumped should have penetrating properties

the main body of fracturing fluid should have a low leakoff rate (at least
during the pumping phase)

that portion of the fluid used to transport proppant should have high
viscosity

With these three idealized characteristics, it should be possible to create a propped


fracture with the required geometry.

A fracture is satisfactory, however, only when we can easily remove the fracturing
fluid from the formation after closure. The fracturing fluid has to allow formation
fluids to flow into and through the induced fracture. Therefore, the final function of
fracturing fluid is to revert to a low-viscosity, non-damaging system that easily
returns to the wellbore and does not leave residue behind.

The presence of polymer residue in the proppant pack is called proppant-pack


damage. The filter cake at the fracture face causes fracture-face damage, and
residue in the polymer invaded zone causes formation damage. Formation damage
can also be caused by non-compatibility of the fracturing fluid or even its filtrate with
the matrix (i.e., clay swelling). In principle, we should keep in mind all the possible
types of damage that may occur during treatment; but in practice, the relative
importance of these elements varies, depending on the formation permeability. For
low and medium permeability fracturing, the major consideration should be to avoid
proppant pack damage. Due to the large area of fracture faces, the other damage
constituents are of less significance. In high permeability formations, however, the
fracture face and formation damage may become critical.

Fluid-Loss Control
Fluid loss results from several sub-processes that occur in the different zones shown
in Figure 1 (Fluid loss as a composed effect ).
Figure 1

Polymer-based fluids continually build up a filter-cake at the fracture face. The filter
cake represents a resistance to flow. The wider it is, the more pressure drop is used
up by the filter cake, and the less is left for driving the fluid into the formation. In the
near vicinity of the fracture face, we distinguish polymer-invaded zone and filtrate
zone. In the polymer-invaded zone, the flow is controlled by the non-Newtonian
polymer rheology. In the filtrate zone, the flow is Newtonian and the viscosity is
effectively that of water. Finally, there is the bulk of the reservoir containing the
original formation fluids. During the leakoff process, the reservoir has to accept fluid;
consequently, it reacts as any reservoir to injection. The reservoir response may be
complicated (transient linear flow, transient pseudo-radial flow; infinite acting or
boundary dominated) and might significantly effect the overall leakoff process.

It is important to understand that the engineering of the fracturing fluid can affect
the sub-processes associated with the filter cake and the polymer-invaded zone, but
it may have little or no influence on the other two sub-processes. Distributing a major
portion of treatment costs to fluid additives might be questionable if the overall
leakoff behavior is mainly controlled by the formation and remains unaffected by the
additives.

Fluid efficiency can be improved by adding gelling agents, special fluid-loss additives,
and specially formulated fluid systems. Viscosifiers added to base fluids increase the
fluid viscosity, thereby increasing pressure drop both through the filter cake and the
polymer invaded zone. Insoluble or slowly soluble fluid-loss additives creating a thin
skin of additional filter cake can also reduce the leakoff rate.
The terms that describe a given fluid’s leakoff characteristics are referred to as
fracturing fluid coefficients. These coefficients are expressed in the same units as the
overall leakoff coefficient, that is, length divided by square root of time. The three
coefficients commonly considered are viscosity controlled, wall-building, and
compressibility controlled fluid loss. For some time, it was believed that the overall
leakoff coefficient could be estimated if these three coefficients (measured in the
laboratory and calculated from formation permeability) and a typical pressure
difference were available. The rule for combining the individual coefficients was
reciprocal averaging or some variant thereof. Some sophisticated fracture design
programs do the calculation from layer to layer and use areal averaging to obtain the
overall leakoff coefficient.

A more practical approach is to consider the laboratory measurements ( Figure 2 ,


Laboratory determination of wall-building fluid loss coefficients) as a source of useful
information helping the fracturing engineer to compare fluid performance relative to
each other.

Figure 2

As far as the quantitative determination of the overall leakoff coefficient is


considered, prediction from laboratory measurements is of limited value. Instead, a
fracture calibration treatment (minifrac) should be carried out to obtain the overall
leakoff coefficient.

Rheology
Fluid viscosity affects a fluid system’s in many ways. The rate of fluid leakoff, the
solid transporting capability of the fluid, the frictional losses in the tubulars and the
cleanup behavior are all strongly related to the rheological properties of the fluid
system. Modern fluid systems can meet multiple goals; the sufficient viscosity to
carry the proppant is only one of them.

Fracturing fluids typically exhibit non-Newtonian rheological behavior. This means


that instead of using a single viscosity value we have to specify a whole curve to
describe their flow behavior. The flow curve, relating the shear stress to shear rate is
most often described by a Power-Law model. This model is flexible enough to reflect
the key element: these fluids gradually loose their viscosity if the flow is more intense
(i.e., they are shear thinning fluids).

A rotating viscometer, such as a Fann Model 50, is used to measure the viscosity
exhibited at different shear rates. Then the Power Law rheological model

v (1)

is applied to determine the fluid’s consistency index (K') and flow behavior index (n').
In Equation 1, is the shear stress and  is the shear rate. Shear rate has a dimension
of reciprocal time and the dimension of shear stress is similar to pressure. The flow
behavior index is dimensionless, but the consistency index has a rather odd
dimension, depending on the value of n'. In the SI system, K' has units of PaSn' while
in field units it is expressed in lbf-secn'/ft2.

The apparent viscosity is defined as the ratio of shear stress to shear rate. In the case
of a Power Law fluid ( Figure 1 ,

Figure 1

Power Law flow curve), this reduces to:

a= K(ddtn’-1 (2)


Apparent viscosity varies with the shear rate. Values for the apparent viscosity of the
fluid at shear rates corresponding to Fann Viscometer measurements at 100 rpm and
300 rpm (shear rates of 170 sec-1 and 511 sec-1, respectively) are typically reported
and used by the industry. Their ready availability makes them useful for comparing
various fluids. Most printed references to apparent viscosity are at one of these shear
rates. These apparent viscosities are frequently taken to be representative of the
viscosity of a fluid in an open fracture (170 sec-1 for 100 rpm on the Fann Viscometer)
and in the tubulars (511 sec-1 for 300 rpm on the Fann Viscometer).

Of particular importance in hydraulic fracturing are the laminar flow regimes in two
limiting geometries ( Figure 2, Idealized cross sections for flow in fracture).

Figure 2

Slot flow occurs in a channel of rectangular cross section with an extremely large
aspect ratio (ratio of the longer side to the shorter side). Limiting ellipsoid flow occurs
in an elliptic cross section with an extremely large aspect ratio. The former
corresponds to the KGD geometry and the latter to the PKN geometry.

If the fluid is Newtonian, single expressions are available to calculate pressure loss in
terms of the viscosity. If the fluid is non-Newtonian, an equivalent viscosity can be
determined which substituted into the Newtonian expression gives the correct
pressure drop. The equivalent viscosity is not the same as the apparent viscosity,
because it is related to the particular flow geometry and reflects the specific shear
rate profile in that geometry. The solutions commonly used in hydraulic fracturing
calculations are summarized below, where uavg denotes volume flow rate divided by
cross sectional area and w0 stand for the maximum width of the elliptical cross
section (for details see Valk? and Economides, 1995).

Pressure Drop and Equivalent viscosity

Newtonian Model:

Slot Flow:

Flow in Limiting Elliptical Cross Section:

Power Law Model:

Slot Flow:

Flow in Limiting Elliptical Cross Section:

In calculating the friction pressure drop through the tubular goods, the well-known
turbulent flow correlations are less useful for fracturing fluids and special relations
have to be applied because of the drag reduction phenomena caused by the long
polymer chains.

Proppant-Carrying Ability and Friction Loss


Fluid viscosity is of paramount importance in proppant transport. We can estimate
the settling velocity of an individual particle in a Power Law fluid (if the viscosity is
large enough to ensure laminar settling) by the following generalization of Stokes'
law:
(1)

where dp is the proppant diameter, and (p-f) is the density difference between the
proppant and the fluid. Such a procedure provides only order-of-magnitude accuracy,
because the settling shear rate usually falls outside the validity range of the Power
Law model. Still, it is often enough to answer the main question:
Is the settling velocity low enough to consider the fluid to be of perfect
transport type?
If the answer to this question is No, the fluid is of drop-out type.

When a low viscosity drop-out type fluid is used, the proppant continually settles
toward the bottom of the fracture as the fluid moves away ( Figure 1, Proppant
transport by a drop-out type fluid).

Figure 1

A bed of proppant is then deposited on the bottom of the fracture, gradually building
up in a dune-type depositional pattern. The pack width is equal to the dynamic
hydraulic width. The pack height is determined by the critical velocity required to
move proppant laterally over the top of the proppant pack. The propped length is
usually much less than the dynamically created length. The pack height continues to
increase after pumping has stopped, since those proppants still in suspension at that
point continue to settle. Settling continues until the fracture has closed on the
proppant. The resulting geometry of the proppant bed is difficult to predict, in fact
most of the proppant may be placed out of the pay layer.

When the more viscous sand-transport type of fluid is used, proppant settling is
negligible and the total created fracture height is propped. After closure, the width of
the proppant pack is more uniformly distributed, both in the vertical and lateral
directions, than it is in the case of a drop-out type fluid.

While thicker fluids generally have better proppant transport capability, a


consideration of other important fluid functions suggests using only as thick a fluid as
is absolutely necessary for reliable solid transport.

Friction Loss
Although high viscosity is generally desirable in a fracturing fluid, it can result in
extreme friction loss when pumping fluid through the wellbore tubulars. Fortunately,
this friction loss is less tightly related to viscosity in the turbulent flow regime than in
the laminar regime, especially for polymer-based fracturing fluids. The long linear
polymer chains have a tendency to suppress turbulence; fracturing fluids therefore
exhibit the so called drag reduction phenomenon and have less friction pressure at
high rates than predicted from correlations valid for Newtonian fluids. Nevertheless,
the friction losses in the tubulars may be responsible for a significant part of the cost
of the treatment because pumping costs are calculated from the horsepower
requirements.

Pressure/Temperature/pH Limitations
Most fracturing fluids are temperature-sensitive. We can overcome this sensitivity to
a certain extent by using special additives (such as temperature stabilizers) to delay
viscosity degradation at high temperatures, or by using additional breakers when
treating low-temperature wells. For very small treatments, the exposure to high
temperature may not last long enough for fluid degradation to become a serious
concern.

Other ways of overcoming temperature sensitivity include pumping larger fluid


volumes to cool down the fracture face, or pumping faster to reduce the exposure
time. An inexpensive cool-down prepad can be pumped ahead of a high-efficiency
fracturing fluid to better utilize the capabilities of the more expensive fluid. High-
temperature fluid systems and special cooldown techniques, which permit the
effective fracture stimulation of wells having a static bottomhole temperature in
excess of 400 oF, are also available.

It can be equally challenging to design fluid systems for use in shallow, low-
temperature wells. Conventional breakers used in crosslinked fluids have to reach a
certain triggering temperature before they can start to act. Thus, shallow wells
typically call for linear (non-crosslinked) fluids that are classified as drop-out type
fluids.

Crosslinked gels are especially sensitive to the pH and hence the chemical nature of
the rock. Therefore, the formation water composition should be taken into
consideration.
Types of Fluids
There are many different types of fluids used in fracture stimulation. Early fracture
treatments almost exclusively used crude oils or special refined oils to ensure
complete compatibility with the reservoir. Water-base systems, the safest and easiest
to use, are currently the most common. There are still several applications for oil-
base and other fluids, with the principal considerations being their compatibility with
the reservoir and their local availability.

Water-Base Fluids

Water may be used in a wide range of formation types, and over a wide range of
temperatures and pressures. It is generally available at low cost, and it is relatively
easy to modify its fluid properties for an additional moderate expense. We can
increase its viscosity, for example, by adding gelling agents. Some of the gelling
agents used as viscosifiers are

natural guar gum (Guar)

hydroxypropyl guar (HPG)

hydroxyethyl cellulose (HEC)

carboxymethyl hydroxyethyl cellulose (CMHEC)

Guar gum has been available in several forms for many years, and is still one of the
most commonly used viscosifiers. It is processed from the commercially grown guar
bean. During the rather crude refining process, some of the insoluble husk is ground
and mixed in with the desired end product. This insoluble material, accounting for
about 9-13 percent of the total solids content, serves to supplement the fluid-loss
additives, thereby improving the frac fluid’s efficiency. There is some concern,
however, regarding the possible detrimental effect of these insoluble materials on
proppant conductivity. Ongoing studies are attempting to accurately quantify these
negative effects. In the meantime, many design engineers have requested the use of
materials containing fewer insoluble solids.

The desire to avoid solids-related problems has spurred the development of such
alternative gelling agents as Hydroxypropyl guar (HPG). This manufactured material
has a total insoluble solids content accounting for less than 3 percent of its total
weight. The viscosity development and friction loss properties of guar and HPG are
about the same. Most water-base fracturing fluid systems use one of these two
viscosifiers.

Cleaner viscosifiers, namely hydroxyethyl cellulose (HEC) and carboxymethyl


hydroxyethyl cellulose (CMHEC), contain less than 1% insoluble materials. They are
used infrequently, however, because of certain temperature limitations, higher cost,
and reduced fluid efficiency due to the lack of supplemental fluid-loss control that
insoluble materials usually provide. These "non wall-building" fluids have been
extensively used in high-permeability formations where fluid induced formation
damage has been of extreme concern. It is questionable, however, as to whether
these fluids are less damaging, because the polymer penetration distance is larger
than for wall-building fluids.

A linear gel (i.e., one that does not incorporate crosslinking chemistry) uses 10 to 100
pounds of gelling agent per 1000 gallons of water, with the usual concentration level
ranging from 25 to 60 pounds per 1000 gallons. A typical HPG gel of 40 lb m/1000 gal
concentration contains 0.48 percent-by-weight of polymer. This concentration
provides enough viscosity for the fluid to carry proppants through the surface
equipment and tubular goods when pumped at a nominal injection rate, although it is
still classified as a drop-out type fluid in terms of bottomhole performance.

Crosslinked gels use similar polymer loadings, but develop considerably higher
viscosities due to intermolecular bonding and the resulting formation of a complex,
three-dimensional network of polymer chains. Through crosslinking, we can increase
a fluid’s viscosity by several orders of magnitude while still maintaining its
"pumpability." Crosslinking agents include Borates and heavy metals such as
zirconium and titanium. The crosslinker is normally added to the slurry after the base
liquid has been blended with the proppant. Sometimes a delaying mechanism is
incorporated using a crosslinking activator to prevent the crosslinking reaction from
taking place until the fluid has reached the fracture. Most crosslinked fluids are
extremely temperature, pH and shear-sensitive. The application of these sensitive
crosslinking and associated delaying agents requires a precise control of the pH,
accurate prediction of the temperature changes and a knowledge of the shear
history.

Shear sensitivity might be reversible or irreversible. If it is irreversible, the


deterioration of viscosity due to shear in the tubulars will be final. For such fluids,
delayed crosslinking might be advantageous (i.e., the crosslinking process takes
place when the fluid is already in the fracture). Table 1 lists some of the variety of
crosslinked fluids in current use.

Crosslinker Gelling pH range Temperature


Agent range oF

B, non- guar, HPG 8-12 70-300


delayed

B, delayed guar, HPG 8-12 70-300

Zr, delayed guar 7-10 150-300

Zr, delayed guar 5-8 70-250

Zr, delayed CMHPG, HPG 9-11 200-400

Zr-a, delayed CMHPG 3-6 70-275

Ti, non- guar, HPG,


delayed

CMHPG 7-9 100-325

Ti, delayed guar, HPG,


CMHPG 7-9 100-325

Al, delayed CMHPG 4-6 70-175

Sb, non- guar, HPG 3-6 60-120


delayed

Table 1 Crosslinked fluid systems

Oil-Base Fluids

Lease crude oil is still used for some fracturing treatments because of its relatively
low cost and compatibility. It is a very inefficient proppant transport medium and has
poor fluid-loss control. But crude oil performance, like that of water, can be improved
with the use of additives. Fluid-loss additives can reduce the leakoff rate to
reasonable values, and new generation viscosifiers allow proppant transport
capabilities comparable with those of crosslinked water-based gels. The friction loss
of gelled oil is much lower than that of un-gelled oil, in some cases approaching that
of gelled water. Still, surface treating pressures for oils generally remain higher
because of the lower hydrostatic pressure of an oil column. Lease oil and gelled oils
are used primarily in formations that are extremely sensitive to water. Occasionally,
Aluminum-phosphate-esters are used in oil-base fluids as viscosifiers and friction loss
reducers.

Acid-Base Fluids

In limestones or dolomitic formations, acids are sometimes used in conjunction with


fracturing fluids to etch flow channels on the formation face. The resulting fracture
conductivity is quite high, as it is proportional to the width of the etched fracture
raised to the third power.

Because of the rapid spending rate of hydrochloric acid (especially at elevated


temperatures), the use of acid-base fluids is limited to treatments in which only
nominal fracture penetration is required, or to treatments on wells with relatively cool
bottomhole temperatures. We can extend this technique’s applicability by using acids
that have retarded spending rates, but the comparative cost of acid versus proppants
further limits its potential economic benefit.

Emulsions

Mixtures of oil and an aqueous material (either water or acid) are sometimes
emulsified and used as fracturing fluids. One such system, commonly referred to as
K-l emulsion, consists of two parts crude oil emulsified in one part of water. This
system is an economical alternative, particularly when the cost of crude oil is low.
The high viscosity of an emulsion creates wider fractures than does an aqueous linear
gel, and assists in reducing fluid leakoff and in transporting the proppants. Emulsions
are especially effective in controlling fluid loss because the fluid that leaks off from
the fracture is a multiphase mixture. The relative permeability to a multiphase
system is always lower than to either single phase. As discussed earlier, mixtures
containing as little as 5% by volume of a second discrete fluid phase have been
effective in limiting leakoff.
In order to minimize friction losses, those emulsions having water as the outside
phase are preferred, although emulsifiers that allow oil to form as the outside phase
of the emulsion are sometimes used for special situations such as in preparing a
retarded (liquid-encapsulated) acid system.

Gas/Foam Fluids

Specialized emulsions using nitrogen or carbon dioxide gas as the inner phase of an
aqueous mixture have become commercially available in recent years. These
emulsified foams typically contain 70 to 90 percent gas at bottomhole fracturing
conditions. This large volume of gas expands even more during cleanup to
supplement the reservoir energy and help with the recovery of injected fluids. The
high viscosity of the foam fluid allows it to transport proppants very efficiently and is
especially beneficial in reducing fluid loss. The multiphase composition of the leakoff
fluid satisfactorily improves the fluid efficiency. The use of foams is especially
effective in highly water-sensitive gas reservoirs where the use of oil is impractical.
The relatively small volume of water (or, in some cases, oil) included in the foam,
coupled with the normally rapid fluid recovery rate, minimizes the detrimental effect
of using water in a water-sensitive reservoir.

Due to the large volumes of expensive gas required at high pressures and the
exceptionally high horsepower required to pump this low-density fluid, this system is
generally not considered for high pressure or deep reservoirs. Exact limitations will-
depend greatly on the relative availability and cost of nitrogen and carbon dioxide,
but the use of high-quality foam as a fracturing fluid is generally limited to wells
shallower than 10,000 ft.

Fracturing Additives
Early fracturing treatments used petroleum-base fluids exclusively, primarily to
ensure the basic fluid functions and compatibility with the reservoir rock and fluids. In
addition to considering reservoir/fluid compatibility, the design engineer must also
that all essential additives are compatible with each other. The highly complex fluid
systems typically used today—especially the crosslinked fracturing fluids—are very
sensitive to even minute changes in additive concentrations. Typical additives include

bacteria control agents

breakers

clay-stabilizing agents

demulsifying agents

dispersing agents

fluid loss additives

foaming agents

friction loss reducers


gypsum inhibitors

nitrogen and carbon dioxide gases

scale inhibitors

sequestering agents

sludge inhibitors

surfactants

temperature-stabilizing agents

water blockage-control agents

Fluid-Loss Additives

Extraneous solids have typically been added to fracturing fluids to reduce the amount
of leakoff into the formation and thereby improve fluid efficiency. The choice of
materials for this purpose has been quite extensive, with the primary requirement
being a mixture of widely ranged particle sizes that will form a relatively impervious
filter cake over the formation pores when the first amount of fluid leaks from the
fracture void into the matrix. The material most commonly used in fracturing
competent sandstone reservoirs is finely ground silica flour.

Larger-size silica particles (100/170 mesh, which is commonly referred to as "100


mesh") have proven quite effective in reducing fluid leakoff into natural fractures.
Other materials are also in use.

Breakers

Chemicals—usually enzymes, oxidizing agents or reducing agents—attack the


viscosifiers in the fracturing fluid, breaking the molecular chains and reducing the
fluid viscosity. This reduced viscosity eases the return of the fluid trapped in the
reservoir pores after the treatment and allows faster and more complete recovery of
the treatment fluids. The effectiveness of chemical breakers depends on their
concentration and temperature. The amount of breaker used in a large frac treatment
is sometimes tapered so as to provide longer stability in the leading portion and a
more rapid break in the last volume of fluid pumped. Breakers may start acting at the
moment of addition, or they may be delayed in action or triggered by another factor
such as temperature.

Other Additives

Potassium chloride (KCl) is commonly used as a special-function additive. This low-


cost salt is generally added at concentrations of about 2% by weight to prepare
artificial formation brine. It prevents the water from adversely affecting the clay
particles present in the formation, and ensures compatibility with the normally saline
formation fluids. In addition, it generally does not interfere with the performance of
other additives. In some cases additional clay stabilizers might be necessary.
Anti-foamers are sometimes required to prevent handling problems at the surface
during the pumping operation. High-rate fluid transfer frequently causes foaming,
particularly if surfactants are used in the fluid. Foaming is more severe if brines are
used instead of fresh water.

Sometimes colonies of sulfate-reducing bacteria start to grow in oil reservoirs as a


result of introducing untreated water into the reservoir during a workover operation.
The use of recycled fracturing fluid containing enzyme viscosity breakers aggravates
this situation. In order to cure and/or prevent this problem, small amounts of bacteria
control agents (biocides) are often included in the fluid.

Formation fines or minute particles such as fluid-loss additives tend to initiate and
stabilize emulsions. Demulsifiers may be required to prevent emulsions from forming
between frac fluids and the formation fluids.

Surfactants are used to reduce the surface and interstitial tension of fluids used in
completion and workover operations. Their use reduces the amount of reservoir
energy expended in removing the treatment fluids from the formation. They also act
as friction pressure drop reduction agents.

Вам также может понравиться