Sei sulla pagina 1di 11

SPE 84310

Avoiding Proppant Flowback in Tight-Gas Completions with Improved Fracture Design


Javier M. Canon, Texas A&M U., Diego J. Romero, El Paso Production Co., Tai T. Pham, El Paso Production Co., Peter
P. Valko,Texas A&M U.
Copyright 2003, Society of Petroleum Engineers Inc.
This paper was prepared for presentation at the SPE Annual Technical Conference and
Exhibition held in Denver, Colorado, U.S.A., 5 8 October 2003.
This paper was selected for presentation by an SPE Program Committee following review of
information contained in an abstract submitted by the author(s). Contents of the paper, as
presented, have not been reviewed by the Society of Petroleum Engineers and are subject to
correction by the author(s). The material, as presented, does not necessarily reflect any
position of the Society of Petroleum Engineers, its officers, or members. Papers presented at
SPE meetings are subject to publication review by Editorial Committees of the Society of
Petroleum Engineers. Electronic reproduction, distribution, or storage of any part of this paper
for commercial purposes without the written consent of the Society of Petroleum Engineers is
prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300
words; illustrations may not be copied. The abstract must contain conspicuous
acknowledgment of where and by whom the paper was presented. Write Librarian, SPE, P.O.
Box 833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.

Abstract
In this paper we describe the possibility and benefits of
incorporating the stability analysis of the proppant pack during
the design stage of hydraulic fracturing treatments.
After the well is treated and cleaned up, the flowback of
proppant from a fracture-treated formation is highly
undesirable for several reasons, including possible damage to
the wellhead and flowlines, operational complications, and last
but not least, decrease in well-productivity. The issue has been
studied on an empirical basis and the most important factors
have been determined. Nevertheless, qualitative models
suggested so far seem to work only under limited conditions
and currently there is no clear methodology for predicting the
occurrence of proppant flowback while designing
the treatment.
In this work we review previously suggested prediction
methods and analyze the proppant flowback patterns
experienced in 24 South-Texas tight-gas completions. As a
result of the study we conclude that the predictive power of
the available models is not satisfactory and the implications
for "proppant stability control agents" are not based on
convincing evidence. A new semi-mechanistic model is
proposed that shows reasonable agreement with both
laboratory and field data. A methodology is suggested to
incorporate the proppant flowback prediction at the fracture
design stage. The suggested methodology is based on the
concept of "minimum necessary departure from optimality to
satisfy technical constraints", in this case the constraint being
to keep the likelihood of proppant flowback under a certain
threshold.
Introduction
Since the late 1940s, there have been more than a million
fracturing treatments performed in the United States1. In a
regular fracturing operation, some flowback of proppant

occurs right after the treatment is completed. This stage is


referred as the clean up phase and is unavoidable since
under-displaced proppant will remain in the wellbore after the
operation. In this stage, personnel and equipment are usually
still in place and any produced solids can be easily handled. In
contrast, flowback during the production phase of a fractured
well can undermine the potential benefits of the stimulation
treatment and represent operational complications. Along this
paper we will refer to this latter situation as
proppant flowback.
The problems associated with proppant flowback are
the following:
Local loss of fracture conductivity, which generates a
reduction in the potential benefits of a hydraulic
fracturing treatment.
Damage to the equipment. (Consisting in abrasion to
valves, tubing, surface pipelines and other equipment).
The first and most obvious approach in dealing with this
problem consisted in applying operational rules of thumb.
The most common of these techniques focused on maintaining
the production rates below a critical value that was determined
as critical to initiate the flowback of proppant in a specific
well. The evolution of forced closure technique, that is
closing the fracture rapidly in order to trap the proppant grains
in a uniform distribution generating a more stable fracture
has been proven usually necessary but not sufficient. It has
been reported2 that these methods do not work all the time,
partly because their justification ignores the actual
mechanisms that create instability in the proppant pack.
Numerous flowback control additives have been offered by
various service companies3. Unfortunately, there is lack of
evidence that they work under conditions of high formation
closure stresses. The use of resin coating in particular, has
been reported unsufficient in some cases4.
On the other hand, some experimental studies5 have helped
deliniate the mechanisms behind proppant flowback.
However, the available predicting models rely on empirical
correlations6,7,8 and may be less reliable as the application
conditions deviate from the conditions of laboratory
experiments on which the correlations are based. The goal of
this work is to summarize available information and improve
the predictive power of the models.
Factors That Determine Proppant Flowback
Although, currently there is no conclusive method to predict
proppant flowback, there is a consensus regarding the
influence of the most important parameters as listed below:

Width. Probably the most sensitive variable in the proppant


flowback phenomenon, the fracture width is believed to be
most influential factor according to some studies5,9. In 1992
Milton-Tayler et al.5 conducted an experimental study aiming
to establish the relative importance of different variables in
determining the likelihood of proppant flowback. From this
study, it was concluded that there is a narrow band of fracture
width ratios (fracture width over mean diameter of a proppant
particle) over which a proppant pack goes from a stable to an
unstable condition. In addition, fractures wider than 6 grains
were always found to be unstable. This finding was
corroborated by Asgian and Cundall9. However, stable
fractures wider than 6 grains were reported elsewhere10.
Though the general conclusion (e.g. fractures wider than 6
particle grains will be unstable) is a result of neglecting the
interaction of many additional parameters with the width, it
applies amazingly well in many practical applications.
Closure Stress. Several theoretical and experimental studies
have evidenced the importance of closure stress since
19925,6,7. It is currently accepted7 that an increase in closure
stress can improve the friction forces among individual grains
yielding a more stable pack. However, it has also been
reported that excessive closure stress can be responsible for
causing proppant flowback7. In this latter case, (e.g. elevated
closure stress) some proppant grains start crushing as the
closure stress approaches the nominal strength of the proppant
material. As result of losing mechanical integrity, the frictioninduced structure throughout the granular pack is lost and
proppant flowback starts.
Drag Forces. The drag forces relate directly to the pressure
drop created when only fluids are moving through the
proppant pack. The pressure drop depends on the viscosity and
velocity of the fluid, and on the permeability of the proppant
pack. Thick fluids will create a larger destabilizing force than
less viscous ones. It has also been shown7 that the influence of
the drag forces is dependent on the closure stress experienced
in the fracture. Basically, the cases where closure stress is
extremely large or extremely low, will both tend to be unstable
and a limited drag force will be enough to destabilize the
proppant pack. In contrast, in cases of intermediate closure
stresses, the drag force is the dominating factor in the proppant
flowback phenomenon.
Proppant Additives. The use of resin coating materials is a
common practice to prevent proppant flowback. These resin
materials help establish a more stable pack in the fracture
under downhole temperature conditions. In addition to the
known limitations (higher cost and the need for specific
curation temperature conditions) the effect of stress
cycling11 has been determined to influence stability in those
cases. A cycle represents producing the well followed by
shutting it in. It is accepted that every fracture with resincoated proppant has a predetermined value of stress cycling it
can tolerate before it starts producing proppant, but obviously
it is not a trivial task to translate actual production history
into theoretically equivalent cycles. Additionally, under high
closure stresses, the bonds formed by the resin coating tend to

SPE 84310

brake easier in a given number of cycles. Therefore, it is not


exceptional that the benefits of the additives are lost4.
In addition to resin coating, many additives are under
development with the goal to improve the stability of the
proppant pack. These include polymers, chopped fibers and
thermoplastics. Some of these products are reported to
increase the friction angle between grains contributing to a
greater stability of the pack. At certain conditions, the critical
flow-rate necessary to initiate granular flow might increase
sufficiently, but the effectiveness of these products seems to
be limited to low-closure stress conditions12,13 (<6,000 psi).
There are no published results regarding the overall effect of
these novel flowback agents under high closure
stress conditions.
Prediction Models
Some models are currently being used in order to predict the
occurrence
of
proppant
flowback
in
practical
applications6,7,8,14. These models can be divided into two
general categories, empirical and theoretical.
Empirical Models
Stimlab Correlation. Since 1996, the Stimlab consortium has
conducted experiments in trying to explain the influence of
different parameters in the occurrence of proppant flowback6.
The results of those studies were used to generate an empirical
correlation. The graphical representation of such correlation is
included in Figure 1. That plot shows the interaction of three
different variables. Net closure stress in the fracture (x axis),
width ratio or number of layers (curve parameter) and critical
fluid velocity (y axis). This critical velocity is a function of the
properties of the produced fluid, as well as of the proppant.
Proppants with larger nominal strengths will be more difficult
to destabilize. The curves in Figure 1, represent the stability
limit for a given condition of width ratio in the fracture. From
this plot it can be seen that fracture stability improves
indefinitely as closure stress increases. Consequently, this
correlation ignores the instability created by the crushing of
grains at elevated values of closure stress. As a consequence,
at values of closure stress larger than 7000 psi, the correlation
is less reliable.
Proppant-Free Wedge Model. This model developed by
Andrews and Kjorholdt7 makes use of the concept of regions
of stability first suggested by the Stimlab Consortium.
However, in this case, the authors did consider the effect of
having partial or even total crushing of the proppant grains at
elevated values of closure stress.
The model is represented in Figure 2. The variables
considered by the model: 1) fracture width, 2) closure stress
and 3) drag force, are translated into three terms: width ratio
(envelope parameter in Fig. 2), closure term (x axis in Fig. 2)
and drag term (y axis in Fig. 2). These two latter terms are
scaled by a reference proppant diameter (equal to 0.0284 inch)
in the following way:

dP d p
F=
dx d ref

.... (1)

SPE 84310

1 d ref
C=
Pc ,net d p

..... (2)

The cubic scaling for the drag term is believed to come


from the body force acting on a proppant due to the
hydraulic gradient. Thus, for the same frictional pressure
gradient, larger proppants will experience a greater
destabilizing force.
The contour lines represent the limit of stability
corresponding to a certain width ratio (Wr). A combination of
drag force and net closure stress (values on the y and x axis
correspondingly) is represented by a point on the plot. If the
point is wrapped by the envelope corresponding to the actual
width ratio, the configuration is stable; otherwise; it is
unstable. The contours of the figure can be obtained from the
following polynomial expression:

Wr,max = 3.2 + 5.51103 C 5.47 105 C 2 + 0.17F ...(3)


+ 1.61102 CF 6.92 103 F 2 5.34 105 C 2 F
The shape of the curves in the model suggests two
different regions. In the first one, where closure stress is low
(C > 1.5 x 10-3), drag forces become the dominant factor.
Here, it can be seen that wide fractures are stable at very low
drag forces. However, a small increase in the flowrate (larger
drag term) can destabilize the pack and generate flowback. As
the closure stress increases (going to the left in the x axis in
Fig. 2), the stability increases due to the improvement of
friction forces, until a point (C ~ 1.0 x 10-3) where it starts to
have the opposite effect. In this second region, mechanical
destabilization is likely to occur. This means that the contact
forces among grains become so large at some localized
regions of the pack, that there is a partial crushing of grains
resulting in a loss of stability.
On the other hand, in the mechanical destabilization area
of the PFW model (left hand side in Fig. 2) the shape of the
envelope curves for Wr = 6 or less, is not convincing. Looking
at the curves, we notice that there could be an unstable
fracture at a low value of drag force that becomes stable by a
small increase in the drag force, keeping all other variables
constant. Evidently, this is an artifact of fitting experimental
data with a polynomial expression that does not stem from
physical laws.
In spite of its evident drawback, and the fact that the
applicability of the PFW model has not been extensively
proven, this simple empirical plot provides an improved
insight into the interaction among the major variables
responsible for the proppant flowback phenomenon.
Theoretical Models. A concept often used in chemical and
mechanical engineering, the minimum fluidization velocity
has been used in various attempts to describe the physical
mechanisms behind the occurrence of proppant flowback8,15.
In a stationary bed of solids, the pressure drop can be
calculated by an expression called the Ergun equation16. If the
fluid velocity is steadily increased, a point is reached at which

the particles no longer remain stationary but fluidize under


the action of liquid or gas. At this point, the porosity of the
pack has also increased, reaching a critical value called
minimum porosity for fluidization, mf. The fluid velocity at
which fluidization begins can be estimated by using the Ergun
equation for packed beds in an extrapolation manner:
(1 mf )( p f ) g = 150
( f / 1488 .16 )v f

[ (d
s

/ 12 ) ]

(1 mf ) 2

mf 3

....(4)

1.75(1 mf )

fv

mf

s ( d p / 12 )

2
f

The procedure for using these models consists of


calculating the fluidization velocity, vf from equation (4). This
can be done in the following way:

vf =

B f + ( B 2f 4 A f C f )
2Af

....... (5)

where:

Af =
Bf =
Cf =

1.75( d p / 12) 2 2f

s mf3 ( f / 1488.16) 2

............ (6)

150( d p / 12) f (1 mf )

f2 mf3 ( f / 1488.16)

............ (7)

g f ( d p / 12) 3 (62.428 SG p f )
( f / 1488.16) 2

..... (8)

The proppant pack is indicated stable if the actual fluid


velocity in the fracture is less than the critical value.
It is seen that this equation does not account for the
influence of closure stress on the walls of the granular
packing. While this approach describes the physical
phenomena when no stress is applied on the faces of the
fracture, it has limited application. Nevertheless, in two
extreme cases it reflects the situation correctly: a) when
closure stress is very low, and grains are indeed cohesionless;
or b) when the width ratio is very large, and hence according
to both the Stimlab correlation and the PFW model, fractures
tend to be unstable regardless of the closure stress.
Semi-Mechanistic Model
It is clear that there are still some discrepancies present in the
models available at present. However, at this point we have
enough elements to integrate different features into one single
formulation, and by doing that we can avoid the shortcomings
found in previous approaches. In this work a new model is
developed with the following features:

SPE 84310

The interaction of net closure stress, drag force and fracture


width is consistently reflected in the shape of the PFW
model. For this reason, the new model will be based on the
same concept of regions of stability.
Because the influence of the width ratio is seen to be highly
relevant, the new model has to reflect the results found in
experimental studies5,6 as well as in theoretical studies9 that
wide fractures will almost always tend to be unstable.
Because there is always a minimum velocity necessary to
mobilize the proppant grains out of the fracture, it is clear
that the curves in Fig. 2 (Proppant-Free Wedge) can never
cross the x axis. In reality, they should tend to flatten out at
a minimum value of pressure gradient, which can be
calculated through the Ergun equation (Eq. 4).
The properties of the proppant also have to be included in
the new formulation. The mechanical destabilization region,
described by the PFW model will then be related to the
strength of the proppant material.
In the new model, the stability criterion calculation starts
with the determination of the drag force term:

Fsta

ln( Pc ,net ) a '


+ FFV ..... (9)
= WT exp 0.5
ST

where the factor WT is an exclusive function of the ratio,


Wr: fracture width divided by average proppant grain
diameter :

WT = 1422.5 exp( 1.0483Wr ) ........ (10)


The width factor WT becomes constant at width ratios
larger than 7.
As it is seen in Eq. 9, this term is proportional to Fsta, so
that when WT is larger the proppant-pack configuration tends
to be more stable. It is also clear that WT suffers a drastic
decrease when the width ratio goes from 2 to 7. For higher
values, WT is small and nearly constant. This behavior can be
explained by a numerical technique called Discrete
Element Method9,17.
On the other hand, in equation 9 the term ST is an
exclusive function of the proppant strength:

S T = 3 10 5 S MAX + 0.22368 ......... (11)


where SMAX represents the nominal strength of the
proppant in units of psia.
To be specific, we define the stress at which the
permeability of the proppant is reduced to 15% of its nonstressed nominal value. In Table 1 we show some typical SMAX
values. More accurate values can be easily deduced from the
standard proppant characterization information provided by
most proppant manufacturers.

The term FFV of equation 9 represents the minimum


pressure gradient for mobilization of loose grains. It correlates
the drag force induced by the production fluids with the
density and size of the proppant grain. Mobilization of a pack
of loose grains starts when the drag force induced by the fluid
flow overcomes the gravity force acting on a grain. As it was
mentioned earlier, this mechanism is described by the Ergun
equation. This equation was solved for the fluid velocity vf.
Once this variable was obtained, the minimum pressure
gradient for fluidization of loose grains, FFV could be
determined from Darcys law in the following way:

FFV = 1.365 10 7

vf f
kf

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

Once the maximum stable pressure gradient, Fsta is


determined from equation 9, this value can be compared to the
actual pressure gradient in the fracture. If the actual pressure
gradient is smaller than the maximum stable value, the
fracture is stable; otherwise, it will be likely to show
proppant flowback.
In Figs. 3 and 4 we show the envelope curves representing
the different stability regions for two types of proppant. Each
curve corresponds to one single width ratio (or number of
layers). The way this plot can be used is the following:
First, the net closure stress acting on the fracture is
estimated and the pressure gradient of the fluid in that system
is calculated.
With this information, a point is located on the plot. If the
point falls inside the envelope curve corresponding to the
actual fracture width ratio, the fracture will be stable. In
contrast, if it falls outside this region, proppant flowback will
be likely to occur.

Analysis of Field Cases


Post Facto Stability Analysis. The data used for this
analysis corresponds to 10 fractured wells in a South-Texas
field operated by El Paso Production Co. Most of these wells
have numerous fractured intervals, so that the total of propped
fractures analyzed is 24. These wells produce mainly gas, and
are characterized by low reservoir permeabilities (0.5-0.05
md) and high temperatures (250F to 350F). Those
completions were identified to suffer from proppant flowback
of variable severity relative to the proppant volume injected.
Our analysis procedure consisted of using three prediction
models (Stimlab correlation, Proppant-Free Wedge Model and
Semi-Mechanistic Model) in order to determine the stability
condition of the wells and compare such results to the relative
amount of proppant actually flowed back. The objective of the
analysis was to establish if proppant flowback is possible to
predict and consequently influence during the design of
the treatment.
General properties of the proppant used in the treatments,
as well as the dimensions of the fractures calculated for this
analysis are included in Tables 2 and 3.

SPE 84310

The results of the stability analysis are shown in Figs. 5 to


7. In those plots, we show a stability variable, which is
obtained by dividing the actual fracture width over the stable
width according to each one of the models. The stability
variable is shown with a solid line. Values of this variable
larger than 1.0 will correspond to unstable configurations,
likely to cause proppant flowback. In the same plots a dashed
line reflects the actual percentage of proppant flowed-back
relative to the total volume injected.
From the figures it is clear that some completions (i.e.
Well 4 Stg. 2, Well 5 Stg. 2) are highly unstable according to
all prediction models. These same wells produced the largest
relative amounts of proppant. On the other hand, for the wells
that reported the lowest relative volumes flowed-back, (i.e.
Well 9 Stg. 4, Well 6 Stg. 1) all the models predicted
stable configurations.
Another important phenomenon present in these field cases
is non-Darcy flow in the fracture. Non-Darcy flow creates
higher frictional pressure gradients in the fracture, increasing
the actual value of the term, F. Therefore, non-Darcy flow
effects are important, when the contour lines of the stability
region tend to be more horizontal than vertical in the plot of
the Semi-Mechanistic Model, (Figs 3 and 4).
An important observation of the stability analysis is that
most of the actually unstable fractures will not achieve
stability even if we reduce the production rate (and
consequently the pressure gradient) to a minimal value. In
those cases, stability cannot be achieved by operational means
(limiting production rate) but only by design means (reducing
the targeted fracture width).
In addition, we found that the most unstable fractures had
width ratios larger than 6. This again, corroborates the initial
hypothesis that fractures wider than 6 particle diameters are
inherently unstable. According to our semi-mechanistic
model, the use of a high strength proppant does not alter this
behavior. The advantage of the semi-mechanistic model
proposed in this work is the clearer description of the
mechanical destabilization region.
While results from the Stimlab Consortium correlation and
Semi-Mechanistic Model may be somewhat more accurate,
one can only reinforce the first empirical observation
documented by Milton-Tayler et al.5: Proppant flowback
becomes a severe problem in wide fractures. It seems that the
issue of applying the right correlation is less important
because of the inherent uncertainty in the estimation of
fracture dimensions. Virtually any reasonable criterion would,
however, improve highly the proppant flowback situation if
considered at the right time: during the design of
the treatment.
Effect on Well Productivity. In addition to the earlier
mentioned analysis, we performed a second series of
calculations for each one of the completions, in which we
evaluated the detrimental effect of proppant flowback on the
productivity of the treated wells.
It is known that the production of proppant has the effect
of creating an additional phenomenon called choked-fracture
skin18. This effect refers to the formation of a damaged zone in
the fracture near the wellbore after proppant flowback occurs.
In this zone, the fracture presents a reduction in width, which

yields a lower conductivity, and ultimately, a decrease in the


overall productivity of the well.
To calculate this skin factor, the length (xck) affected by a
reduction in width as consequence of the flowback of proppant
must be determined. From the mass of proppant produced, we
first calculate the fracture volume lost due to
proppannt flowback:

VPFB =

M PFB
... (13)
62.75(1 pp ) SG p

Then, we assume a final stable width (Wstable) where


flowback stops and considering that the proppant produced
comes from the whole height of the net pay, it is possible to
calculate xck, which represents the choked length in one wing
according to:

xck =

V PFB
..... (14)
2h p (W p ,max Wstable )0.0833

Subsequently, the skin factor created by the flowback of


proppant is calculated using the following expression18

sck =

xck W p ,avg

1 ........ (15)

Wstable

xf

The effect of the created skin is a reduction of the


dimensionless productivity index:

J d ,ck =

1
1
J d ,id

.......... (16)

+ sck

where
Jd,id
denotes the optimum dimensionless
productivity index (without the choke in the fracture)
calculated from the actual proppant mass reaching the
pay layer.
Another dimensionless productivity index, Jd,sta. can be
calculated from the the maximum amount of proppant that
when placed optimally could still satisfy the
stability criterion.
Shown in Figure 8 are, for some of the wells analyzed, the
actual proppant mass injected and the maximum amount of
proppant that could still satisfy the stability criterion
(according to the various models).
For comparison of well productivities on Figs 9-11, we
show the ideal productivity which assumes no proppant
flowback and optimum fracture dimensions, denoted by Jd,id.
On the same figures the choked productivity, Jd,ck is shown
with dark color. A third bar shows Jd,sta, the best productivity
achievable if we design the treatment to avoid proppant
flowback. For brevity, we call this the proposed treatment.
Obviously, those designs considering the stability criteria

SPE 84310

result in less expensive treatments, because the overall amount


of proppant injected is considerable less. Nevertheless, from
Figs. 9 through 11 it is clear that even though there are
differences among the methods, in most cases a lower amount
of proppant would have created a stable fracture and a more
productive well than the actual treatment. We see, for
instance, that in Well 5 Stg. 1 and 2, proppant flowback was
responsible for reducing severly the productivity of the wells.
In those cases, the proposed fractures would have been more
convenient (and would have also saved operational costs
stemming from proppant back-production ).
These results confirm the benefits of including a stability
criterion during the design of a fracturing treatment. For this
purpose, we suggest a methodology (Figure 12) that consists
in modifying the amount of proppant injected, in cases where
proppant flowback is likely to occur according to the stability
models. As a main result of this work, we suggest to modify
the target amount of proppant, that is treatment size. This not
excludes the application of other means, including proppant
flowback agents, and posing operational limit on the
production rate, but seems to be more fundamental.

changes in net closure pressure and drag force. If we assume


that the width ratio in the choke region of these fractures is
reduced to approximetly 3.5 as a consequence of proppant
flowback, these two completions will be stable at t2 according
to the prediction model.
While the resin coated material used in the fracture might
have helped to delay the flowback of proppant, the pack
ultimately failed. An intermittent production of proppant then
resulted in a reduced width in the choke region. The reduced
width brought back the well into the stable region. It is
anticipated that an additional decrease in flowing bottomhole
pressure will move the point again into the unstable region.
Indeed, similar cyclic proppant flowback phenomena have
been observed in several neighboring wells.

Proppant Flowback with Time. In order to analyze the


flowback mechanism in detail, we focused on a specific well.
Well 11 cleaned up producing very little amounts of
proppant during its several weeks on line during which it was
producing commingled from two formations that were
independently treated (Stg. 1 & 2). It actually started to flow
proppant after a 20 day proppant free period, when the FWHP
was reduced to 3,200 psi. The proppant pack stability was
evaluated using the semi-mechanistic model right before the
well started flowing proppant back to the surface. The results
of this analysis correspond to to the time point t1
(approximately 19 days) in Table 4 and Figure 13.
The proppant used was resin-coated bauxite. From the
calculations using the semi-mechanistic model (Table 4 and
Fig. 13), it was observed that Stg. 1 was unstable and it would
flow proppant at the given conditions. However, the well did
not produce significant amounts of proppant during the
cleanup. This behavior could be attributed to the fact that the
resin was holding the proppant grains together preventing the
flowback at the initial conditions.
During the production period before the proppant flowedback, the well head pressure decreased from 6000 to 3300 psi,
which corresponds to a similar increase in the net closure
stress. It has been observed in laboratory4 that the resin
detaches from the proppant at net closure of 5,500 to 6,500
psi. Interestingly, the laboratory data are in good accordance
with the fact, that the net closure stress on proppant at the
moment of failure was approximately 5700 psi .
Also, note that this pressure corresponds to 4,800-5,200 psi
bottom hole flowing pressure, which is the apparent dew point
of the reservoir fluid. The presence of condensate causes a
large increase in the drag force term (two-phase non-darcy
flow) and hence it is not surprising, that the proppant pack
became unstable. It is anticipated that proppant will flow back
until a stable width ratio is reached in the vicinity of the well.
In Table 4 and Fig. 13 we depict this situation (time point = t2
approximately 21 days). In the semi-mechanistic plot (Fig. 13)
we see the displacement of the points as a result of the

2. It is suggested to include a proppant flowback criterion


into the design procedure of hydraulic fracturing treatments.
An optimum treatment design should target a stable proppant
pack, even if might lead to a reduction of the treatment size
(with respect to original expectations.) The ultimate goal is to
avoid flowback and maximize the long-term productivity of
the treated well.

Conclusions
1. Not considering the stability of the proppant pack during the
design of hydraulic fracturing treatments has translated into
the execution of sub-optimal treatments causing increased
operational costs and ultimately decreasing productivity of the
treated wells.

3. The semi-mechanistic model for predicting proppant


flowback provides a moderate improvement over the currently
available models. The approach accounts for a wide range of
various conditions, including net closure stress, fluid
properties, and proppant characteristics.
4. The semi-mechanistic model is suitable to analyze actual
field cases, and the insight gained is sufficient to explain the
basic observed trends. In general terms, the most unstable
fractures according to the prediction models are the ones with
the largest relative volumes of proppant actually flowed-back.
5. Fractures with width ratios larger than 6 or 7 particles are
consistently found as unstable. This hypothesis, first suggested
by Milton-Tayler, still remains true when more detailed
models are used and the tendency is also corroborated by
field data.
6. Changes in the bottomhole flowing pressure of a well (and
reductions in the fracture width as a consequence of the
flowback of proppant) can create alterations in the stability
condition of a fractured-treated well. For this reason it is
useful to analyze the interaction of these variables in time with
the aid of the semi-mechanistic model.
Nomenclature
a
Af
Bf

: Constant for the semi-mechanistic model = 7.7172


: Term for the calculation of the minimum
fluidization velocity
: Term for the calculation of the minimum
fluidization velocity

SPE 84310

C
Cf

: Closure term in the PFW model, psia


: Term for the calculation of the minimum fluidization
velocity
Cfd
: Dimensionless fracture conductivity
: Mean grain diameter of proppant, in
dp
dref
: Reference diameter, 0.0284 in
dP/dx : Pressure gradient, psi/ft
F
: Drag term, psi/ft
Factual : Actual pressure gradient in the fracture, psi/ft
FFV
: Minimum pressure gradient to destabilize loose proppant
grains, psi/ft
Fsta
: Maximum stable pressure gradient in the semimechanistic model, psi/ft
FBHP : Flowing bottomhole pressure, psia
FWHP : Flowing wellhead pressure, psia
g
: Acceleration of gravity : 32.2 ft/ s2
: Fracture height, ft
hf
hp
: Net pay thickness, ft
: Ideal dimensionless productivity index
Jd,id
Jd,ck
: Dimensionless productivity index after choked
fracture efffect
Jd,sta
: Best achievable dimensionless productivity index
kf
: Proppant pack permeability, md
: Mass of proppant, lbs
MP
MPFB
: Mass of proppant flowed back, lbs
: Net closure pressure, psia
Pc,net
sck
: Choked fracture skin
: Specific gravity of proppant
SGp
: Maximum strength of proppant, psia
SMAX
: Strength term
ST
vf
: Minimum fluidization velocity, ft/s
: Volume of proppant placed in the net pay, ft3
Vfp
Vi
: Volume of proppant injected into pay, ft3
VPFB
: Volume of proppant flowed back, ft3
Wr
: Width ratio
Wr,max : Maximum stable width ratio
: Propped width, in
Wp
Wp,avg : Average propped width, in
Wp,max : Maximum propped width, in
Wstable : Stable width, in
WT
: Width factor in the semi-mechanistic model
: Choked length in one wing, ft
xck
: Fracture half-length, ft
xf
Greek Letter Variables

mf
f
f
p
s
pp

: Mean dense phase voidage at minimum fluidization


: Fluid viscosity, cp
: Fluid density, lb/ft3.
: Particle density, lb/ft3
: Shape factor for granular particles in the Ergun equation
: Proppant pack porosity

SI Metric Conversion Factors


cp 1.0
E-03 = Pas
ft 3.048
E-01 = m
in. 2.54*
E+00 = cm
md 9.869 233E-04 = m2
psi 6.894 757E+00 = kPa
bbl 1.589 873E-01 = m3
*Conversion factor is exact.

References
1. NSI Technologies (eds.), Hydraulic Fracturing, Tulsa, Oklahoma
(1987) 3.
2. Gidley J.L., Holditch S.A., Nierode D.E. and Veath R.W.: Recent
Advances in Hydraulic Fracturing, Monograph Series, SPE,
Richardson, TX (1989) 12, 210-222.
3. Stephenson, C.J., Rickards, A.R., Brannon H.D.: Increased
Resistance to Proppant Flowback by Adding Deformable
Particles to Proppant Packs Tested in Laboratory, paper SPE
56593 presented at the 1999 SPE Annual Technical Conference
and Exhibition, Houston, D.C., 3-6 October.
4. Frac Tech Ltd.: Proppant Flowback Report, 2002.
5. Milton-Tayler, D., Stephenson, C. and Asgian, M.: Factors
Affecting the Stability of Proppant in Propped Fractures: Results
of a Laboratory Study, paper SPE 24821 presented at the 1992
SPE Annual Technical Conference and Exhibition, Washington,
D.C., 4-7 October.
6. Stimlab Consortium Reports, 1996-2002
7. Andrews, J.S. and Kjrholt H.: Rock Mechanical Principles
Help Predict Proppant Flowback from Hydraulic Fractures,
paper SPE 47382 presented at the 1998 SPE/ISRM Eurock
Conference, Trondheim, Norway, 8-10 July.
8. Parker, M., Weaver, J. and Van Batenburg, D.: Understanding
Proppant Flowback, paper SPE 56726 presented at the 1999 SPE
Annual Technical Conference and Exhibition, Houston,
3-6 October.
9. Asgian, M.I. and Cundall, P.A.: The Mechanical Stability of
Propped Hydraulic Fractures, paper SPE 28510 presented at the
1994 SPE Annual Technical Conference and Exhibition, New
Orleans, 25-28 September.
10. Hall, C.D. and Harrisberger, W.H.: Stability of Sand Arches: A
Key to Sand Control, JPT (July 1970) 821-823.
11. Vreeburg, R-J, Roodhart, L.P. and Davies, D.R.: Proppant Back
Production During Hydraulic Fracturing: A New Failure
Mechanism for Resin-Coated Proppants, paper SPE 27382
presented at the 1994 SPE International Symposium on
Formation Damage Control, Lafayette, Louisiana, 7-10 February.
12. Nguyen, P. et al.: Proppant Flowback Control Additives, paper
SPE 36689 presented at the 1996 SPE Annual Technical
Conference and Exhibition, Denver, 6-9 October.
13. Nguyen, P.D., et al.: Surface-Modification System for FractureConductivity Enhancement, paper SPE 48897 presented at the
1998 SPE International Conference and Exhibition, Beijing, 2-6
November.
14. Wang, J., Conway, M. and Barree, R.D.: Bi-Power Law
Correlations for Sediment Transport in Pressure Driven Channel
Flows, Journal of Fluid Mechanics, to appear.
15. Sparlin, D.D. and Hagen, R. W.: Proppant Selection for
Fracturing and Sand Control, World Oil (January 1995) 37-40.

SPE 84310

16. W. L. McCabe and J.C. Smith (eds.), Unit Operations of


Chemical Engineering, McGraw-Hill Book Co. Inc., New York
(1976) 146-169.
17. Canon, J.M.: Predicting Proppant Flowback from FractureStimulated Wells, Master of Science Thesis, Texas A&M
University, 2003.
18. Romero, D.J. Valk, P.P and Economides, M J.: The
optimization of the productivity index and the fracture geometry
of a stimulated well with fracture face and choke skins, SPE
Production and Facilities, (February) 2003, pp 57-63.

Table 1 - Values of Closure Stress for various Proppants at


which the Proppant Pack Permeability is Reduced to 15%
Accepted for this Study
(Sorce: various cataloges of manufacturers)
Proppant

Size,
mesh

SGP

kf
nominal,
md

kf
reduced,
md

Nominal
Strength,
psia

Brady Sand

12/20

2.65

1,000,000

150,000

6,500

Brady Sand

16/30

2.65

300,000

45,000

7,600

Brady Sand

20/40

2.65

300,000

45,000

6,420

Table 2 - Proppant Characteristics for the Wells Incorporated


into the Post Facto Stability Analysis
Proppant
Well

kf , md

Proppant
Type

Well 1 Stg. 1

50,000

Ceramax-P

Mesh
Size

SGp

dp, in

636,877

20/40

3.45

0.0248

Mp, lbs

Well 2

50,000

Ceramax-P

1,102,000

20/40

3.45

0.0260

Well 4 Stg. 1

50,000

SHS Bauxite

668,166

20/40

3.34

0.0248

Well 4 Stg. 2

50,000

SHS Bauxite

211,320

20/40

3.34

0.0248

Well 5 Stg. 1

50,000

SHS Bauxite

535,000

20/40

3.34

0.0248

Well 5 Stg. 2

50,000

SHS Bauxite

400,000

20/40

3.34

0.0248

Well 6 Stg. 1

50,000

469,540

18/30

3.34

0.0306

Well 6 Stg. 2

50,000

544,000

18/30

3.34

0.0306

Well 6 Stg. 3

50,000

263,271

18/30

3.34

0.0306

Well 6 Stg. 4

50,000

190,500

18/30

3.34

0.0306

Well 7 Stg. 1

50,000

587,500

20/40

3.34

0.0248

Well 7 Stg. 2

50,000

381,300

16/30

3.34

0.0351

Well 7 Stg. 3

50,000

284,267

20/40

3.34

0.0248

Well 8 Stg. 1

50,000

359,088

20/40

3.34

0.0248

Well 8 Stg. 3

50,000

530,120

20/40

3.34

0.0248

Sintered
Bauxite
Sintered
Bauxite
Sintered
Bauxite
Sintered
Bauxite
Sintered
Bauxite
Sintered
Bauxite
Sintered
Bauxite
Sintered
Bauxite
Sintered
Bauxite
Sintered
Bauxite

Hickory Sand

12/20

2.65

1,000,000

150,000

6,550

Well 8 Stg. 4

50,000

165,404

20/40

3.34

0.0248

Colorado Silica Sand

16/30

2.65

300,000

45,000

4,800

Well 9 Stg. 1

50,000

Sintered Ball

507,700

16/30

3.34

0.0351

Ohio Sandstone

16/20

2.65

350,000

52,500

13,700

Well 9 Stg. 2

50,000

Sintered Ball

678,550

16/30

3.34

0.0351

Ohio Sandstone

16/30

2.65

350,000

52,500

11,200

Well 9 Stg. 3

50,000

Sintered Ball

640,310

16/30

3.34

0.0351

Ohio Sandstone

20/40

2.65

250,000

37,500

9,038

Well 9 Stg. 4

50,000

Sintered Ball

837,810

16/30

3.34

0.0351

Sinterball

16/20

3.62

360,000

54,000

19,200

Well 10 Stg. 1

50,000

676,500

20/40

3.34

0.0248

Sinterball

16/30

3.62

360,000

54,000

17,800

Sinterball

20/40

3.62

360,000

54,000

18,400

Well 10 Stg. 2

50,000

445,500

20/40

3.34

0.0248

Lt Wt Ceramics

20/40

2.7

360,000

54,000

12,100

IS Ceramics

20/40

3.2

385,000

57,750

14,050

HS Ceramics

20/40

3.5

539,000

80,850

16,200

Sintered
Bauxite
Sintered
Bauxite

Table 3 Results of the Design Calculation


Design Calculation
Well

hf , ft

Vi ,
ft3

hp/hf

Cfd

Jd,opt

Xf , ft

Wp, in

Well 1 Stg. 1

142

2,274

0.901

11.96

1.646

818

0.235

Well 2

354

3,935

0.488

7.968

1.522

835

0.160

Well 4 Stg. 1

200

2,464

0.800

12.96

1.662

689

0.214

Well 4 Stg. 2

70

779

0.800

11.71

1.641

689

0.194

Well 5 Stg. 1

170

1,973

0.906

19.73

1.736

542

0.257

Well 5 Stg. 2

150

1,475

0.800

16.71

1.709

542

0.218

Well 6 Stg. 1

155

1,732

0.929

3.31

1.111

1298

0.103

Well 6 Stg. 2

165

2,006

0.988

3.40

1.129

1336

0.109

Well 6 Stg. 3

100

971

0.880

3.17

1.087

1236

0.094

Well 6 Stg. 4

63

703

0.810

3.31

1.111

1297

0.103

Well 7 Stg. 1

178

2,167

0.882

3.40

1.129

1336

0.109

Well 7 Stg. 2

120

1,406

0.967

3.36

1.121

1319

0.107

Well 7 Stg. 3

85

1,048

0.588

3.42

1.132

1342

0.110

Well 8 Stg. 1

109

1,324

1.000

3.40

1.128

1336

0.109

Well 8 Stg. 3

160

1,955

1.000

3.41

1.130

1338

0.110

Well 8 Stg. 4

55

610

0.945

3.30

1.110

1295

0.103

Well 9 Stg. 1

148

1,873

1.000

3.45

1.138

1354

0.112

Well 9 Stg. 2

204

2,503

0.868

3.41

1.131

1340

0.110

Well 9 Stg. 3

189

2,362

0.714

3.43

1.135

1348

0.111

Well 9 Stg. 4

254

3,090

1.000

3.40

1.129

1336

0.109

Well 10 Stg. 1

170

2,495

1.000

3.63

1.175

1420

0.124

Well 10 Stg. 2

112

1,643

1.000

3.63

1.175

1420

0.124

SPE 84310

Table 4 Results of the Proppant Flowback with Time


Analysis
Well

FBHP,
psia

Wr

Pc,net,
psia

Fsta,
psi/ft

Factual,
psi/ft

t1

Well 11 Stg. 1

4716

8.86

5784

0.49

4.18

t1

Well 11 Stg. 2

4716

5.64

5784

2.00

6.27

t2

Well 11 Stg. 1

3300

3.5*

7200

9.88

6.27

t2

Well 11 Stg. 2

3300

3.5*

7200

9.88

9.40

50

F , psi/ft

Time
Point

20/40 Sand Proppant


60

40

3 Layers

30

4
20

* Assumed Value

10

6
0
0

Fluid Velocity,ft/s

6,000

8,000
P c,net
, psia 10,000

12,000

14,000

16,000

In this area 6
layers stay in
and 7 flow out

4 Layers

0.1

4,000

Fig. 3 - Contour Curves for 20/40 Sand According to the Proposed


Semi-Mechanistic Model.

3 Layers

0.12

2,000

5 Layers

0.08

6 Layers

0.06

20/40 High Strength Ceramics Proppant

7 Layers

60

0.04

50

8 Layers

0.02

F , psi/ft

40

0
0

1,000

2,000

3,000

4,000

5,000

6,000

Closure Stress, psia

7,000

8,000

9,000

3 Layers

30
4

20

Fig. 1 Graphical Representation of the Stimlab Consortium


Correlation (Reference 6).

10
6

0
0

2,000

4,000

6,000

8,000
10,000
P c,net , psia

12,000

14,000

16,000

Fig. 4 - Contour Curves for 20/40 High Strength Ceramics


Proppant According to the Proposed Semi-Mechanistic Model.

16
14
Stability Analysis With Stimlab Correlation
18

4
6

8
7

4
8

0.50%

Fig. 2 - Original Representation of the Proppant Free Wedge Plot


(Reference 7).

Well 1 Stg.1

Well 5 Stg. 1

Well 5 Stg. 2

Well 4 Stg. 1

Well 2

Well 4 Stg. 2

Well 10 Stg. 1

Well 10 Stg. 2

Well 8 Stg.3

Well 7 Stg. 3

Well 8 Stg.1

Well 7 Stg. 1

Well 8 Stg. 4

Well 6 Stg. 2

Well 6 Stg. 1

2.0E-03

-1

Well 6 Stg. 4

1.5E-03

1/C , psia

0.00%
Well 9 Stg. 1

10

Well 9 Stg. 3

1.0E-03

1.00%

Well 9 Stg. 2

5.0E-04

9
0
0.0E+00

1.50%

10

Well 9 Stg. 4

2.00%

12

Well 6 Stg. 3

%Proppant Flowed-Back

14

Well 7 Stg. 2

F , psi/ft

Actual Width / Stable Width

10

2.50%

Stimlab Correlation Results

16

Percentage of Proppant BackProduced

12

Well

Fig. 5 Post Facto Stability Analysis Results According to the


Stimlab Correlation.

10

SPE 84310

Stability Analysis With Semi-Mechanistic Model


4

2.50%

1.8

%Proppant Flowed-Back
2.00%

3
2.5

1.50%

2
1.00%

1.5
1

0.50%
0.5

1.3
1.2

Well 1 Stg.1

Well 5 Stg. 1

Well 5 Stg. 2

Well 4 Stg. 1

Well 2

Well 4 Stg. 2

Well 10 Stg. 1

Well 10 Stg. 2

Well 8 Stg.3

Well 7 Stg. 3

Well 8 Stg.1

Well 7 Stg. 1

Well 8 Stg. 4

Well 6 Stg. 2

Well 6 Stg. 1

Well 6 Stg. 4

Well 9 Stg. 1

Well 9 Stg. 3

Well 9 Stg. 2

Well 9 Stg. 4

Well 6 Stg. 3

1.4

1.1
1.0
Well 1 Stg 1

1.8

2.00%

1.6

1.50%

1.4
1.2

1.00%

1
0.8

0.50%

0.6

Well 5 Stg. 1

Well 1 Stg.1

Well 5 Stg. 2

Well 4 Stg. 1

Well 4 Stg. 2

Well 2

Well 10 Stg. 1

Well 10 Stg. 2

Well 8 Stg.3

Well 7 Stg. 3

Well 7 Stg. 1

Well 8 Stg.1

Well 8 Stg. 4

Well 6 Stg. 2

Well 6 Stg. 1

Well 6 Stg. 4

Well 9 Stg. 1

Well 9 Stg. 3

Well 9 Stg. 2

Well 9 Stg. 4

Well 6 Stg. 3

0.00%
Well 7 Stg. 2

0.4

Jd,ck

Jd,id

Jd,sta

1.7
1.6
1.5
Jd

Actual Width / Stable Width

2
1.8

Percentage of Proppant BackProduced

2.50%

%Proppant Flowed-Back

Well 4 Stg 1 Well 4 Stg 2 Well 5 Stg 1 Well 5 Stg 2

Fig. 9 Comparison of Dimensionless Productivity Indexes


(PFW Model)

Stability Analysis With Proppant-Free Model


Proppant-Free Wedge Model

Well 2

Well

Fig. 6 - Post Facto Stability Analysis Results According to the


Semi-Mechanistic Model.

2.2

Jd,sta

1.5

Well

2.4

Jd,id

1.6

0.00%
Well 7 Stg. 2

Jd,ck

1.7

Jd

Actual Width / Stable Width

3.5

Percentage of Proppant BackProduced

Semi-Mechanistic Model Results

1.4
1.3
1.2
1.1
1.0
Well 1 Stg 1

Well 2

Well

Fig. 7 - Post Facto Stability Analysis Results According to the


Proppant-Free Wedge Model.

Well 4 Stg 1

Well 4 Stg 2

Well

Well 5 Stg 1

Well 5 Stg 2

Fig. 10 - Comparison of Dimensionless Productivity Indexes


(Stimlab Correlation)

Mass of Proppant Injected (lbs)

1,200,000

Actual Mass Injected


Stable Fracture (PFW Model)
Stable Fracture (Semi-Mechanistic Model)
Stable Fracture (Stimlab Correlation)

1,000,000

Jd,ck

Jd,id

Jd,sta

1.7
1.6
1.5
Jd

800,000

1.8

600,000

1.4
1.3

400,000

1.2
200,000

1.1
1.0

0
Well 1 Stg 1

Well 2

Well 4 Stg 1 Well 4 Stg 2


Well

Well 5 Stg 1

Well 5 Stg 2

Fig. 8 Proppant Mass Injected for Actual Cases and Hypothetical


Stable Fractures According to Stability Models

Well 1 Stg 1

Well 2

Well 4 Stg 1

Well 4 Stg 2

Well 5 Stg 1

Well 5 Stg 2

Well

Fig. 11- Comparison of Dimensionless Productivity Indexes


(Semi-Mechanistic Model)

SPE 84310

11

Calculate Optimum Dimensions of Fracture18


Stability Criterion

PFW Model

Stimlab
Correlation

Semi-Mechanistic
Model
Most conservative

Stable Fracture

No

Reduce Proppant Mass

Yes
Implement Design

Fig. 12 - Suggested Methodology for Incorporating a Stable Fracture Criterion into a Hydraulic Fracture
Treatment Design.

Semi-Mechanistic Plot (Well 11 Stg. 1 & 2 at different times)

30
Well 11 Stg. 1 at Time = t1
25

3 Layers
4

Well 11 Stg. 2 at Time = t1

20

F (psi/ft)

Well 11 Stg. 1 at Time = t2

Well 11 Stg. 2 at Time = t2

t1< t2

15

Wr=8.8
10

Wr=5.6

6
7

0
0

2,000

4,000

6,000

8,000

Pc,net (psia)

10,000

12,000

14,000

16,000

Fig. 13 Semi-Mechanistic-type Plot of the Proppant Flowback with Time Analysis

Potrebbero piacerti anche