Sei sulla pagina 1di 10

Journal of Petroleum Science and Engineering 134 (2015) 8796

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


journal homepage: www.elsevier.com/locate/petrol

Rheology of polymer-free foam fracturing uids


M. Gu, K.K. Mohanty n
The University of Texas at Austin, United States

art ic l e i nf o

a b s t r a c t

Article history:
Received 2 July 2013
Received in revised form
15 January 2015
Accepted 18 July 2015
Available online 20 July 2015

Aqueous polymeric solutions are commonly used for fracturing in conventional petroleum reservoirs for
their ability to transport proppants deep into fractures, but such uids are not used for shale reservoirs
because of the plugging of nano-pores of shales by large polymer molecules. Polymer-free aqueous foams
can be used in shale fracturing because of their capacity to suspend proppants and lower water consumption than commonly used water-based fracturing uids. In this work, the rheology of three kinds of
polymer-free foams (A: 0.5 wt% regular anionic surfactant, B: Fluid A 2 wt% glycerol, C: 0.5 wt% viscoelastic surfactant or VES) are studied in a recirculating pipe rheometer up to 155 F and 2000 psi. These
temperature and pressure conditions are lower than those for typical shales of current interest. All three
foams exhibit power-law rheological behavior. The regular surfactant foams (A and B) show shear
thinning behavior at qualities above 60%, non-shear dependent behavior from 50% to 60% and shear
thickening below 50% (due to turbulence). The VES foams show shear thinning at qualities less than 60%.
Temperature lowers the viscosity of foams by decreasing the liquid phase viscosity and the stability of
the bubbles, but the temperature effect is small for foams A and B. Pressure increases foam viscosity; the
increase is higher for higher quality foams. As the pressure increases, the rate of viscosity increase decreases. The aqueous foams A and B are both less viscous than typical guar foams (0.24 wt% polymer),
while the VES foam C has a similar viscosity. The model proposed by Brouwers (2010) for suspensions
agrees the best with the measured data for quality under 60%; the model proposed by Princen and Kiss
(1989) for high quality does not match the experimental data. New correlations have been developed to
describe the aqueous foam rheology as a function of shear rate, quality, and pressure at the parameter
range typical of hydraulic fracturing. The correlations can be incorporated in a fracture propagation simulator to evaluate the foam fracturing efciency.
& 2015 Elsevier B.V. All rights reserved.

Keywords:
Foam
Rheology
Hydraulic fracture
Surfactant
Shale fracturing uid

1. Introduction
Production of oil and gas from shale reservoirs has revitalized
the domestic energy production in US in the last 5 years (King,
2010). The key to production from shales is hydraulic fracturing
where horizontal wells are drilled and multiple (1050) fractures
are created perpendicular to the horizontal wells. These fractures
increase the contact between the very low permeability shale
matrix and the production well. The current normal fracturing
treatment in shale gas involves pumping large volumes of slickwater (water with a small amount of a drag-reducing polymer)
along with proppants (typically well sorted sand) at a high
pumping rate. The slickwater frac uids are effective in creating
large, complex fracture networks at a lower cost. However, if
conventional proppants, like sands are used in slickwater, they
n
Correspondence to: Petroleum and Geosystems Engineering, Mail Stop C0304,
The University of Texas, 200 Dean Keeton, Austin, TX 78712, USA.
Fax: 1 512 471 9605.
E-mail address: mohanty@mail.utexas.edu (K.K. Mohanty).

http://dx.doi.org/10.1016/j.petrol.2015.07.018
0920-4105/& 2015 Elsevier B.V. All rights reserved.

settle down within a small distance away from the wellbore,


leaving a lot of fractured surfaces unpropped or closed after fracturing operations (Brannon and Starks, 2009; Cipolla et al., 2010;
Warpinski et al., 2009).
The proppants can be placed further away from the well bore in
a fracture if a viscous fracturing uid is used. There are usually two
ways to increase the water viscosity. One way is to add polymers,
such as guar and guar derivatives (Gidley et al., 1990). However, in
unconventional shale reservoirs, the large polymer molecules can
plug the small pores of the fracture surface and decrease the gas
ow (Peles et al., 2002). The second way is to use foam-based
fracturing uids (Gidley et al., 1990). The cellular structure of
foams can give a high effective viscosity without plugging up the
shale pores if polymers are avoided. Foams are usually generated
from a base uid made from a surfactant, a polymer stabilizer such
as guar, HPG, and Xanthan gums. Foams made from a surfactant
foamer without polymers are known as polymer-free (or gel-free)
aqueous foams. The foam rheology is a very important property for
fracture-treatment design. It inuences the tubing pressure drops,
pump pressure head, fracture geometry, proppant transport and

88

M. Gu, K.K. Mohanty / Journal of Petroleum Science and Engineering 134 (2015) 8796

Re
V
Q
d

Nomenclature
H

P
m
L

K
Ca

dsv
v
s

viscosity ratio, no unit


Taylor coefcient, no unit
pressure difference, Pa
viscosity, cp
length of testing tube, m
stress, pa
power law index, no unit
shear rate, s  1
consistency index, Nsn/m2
capillary number, no unit
uid density, kg/m3
sauter mean diameter, m
ow rate, m/s
interfacial tension, mN/m

Subscripts
in
ex
g
l
w
wi
wa
a

uid loss to the matrix. It is affected by foam quality, texture,


viscosity of the base aqueous phase, pressure and temperature. A
circulating loop rheometer is widely used to characterize the foam
rheology.
Hutchins (2005) investigated the dynamic stability and rheology of foams containing polymers up to 400 F and 2000 psi by
using a circulating loop foam rheometer. A power law model was
proposed to describe the effective viscosity of foams. Foam stability and bubble texture were affected by the additives, temperature, and pressure. The foams with ne and homogeneous bubble
texture were more viscous and stable than those with coarse and
heterogeneous texture. Bonilla and Shah (2000), Sani et al. (2001),
and Sudhakar (2002, 2003) utilized a similar circulating loop
rheometer to investigate the rheology of Guar foams and Xanthan
foams. HerschelBulkley (HB) model and power law model were
adopted to correlate the viscosity data.
Harris and coworkers (1987, 1989, 1995, 1996) have done a
comprehensive study on polymeric foam rheology. They found
that foam ow behavior can be described by HerschelBulkley
model with the power law index, n, the same as that of the liquid
phase and the consistency index, K, as a function of the liquid
phase consistency index and foam quality (Reidenbach et al.,
1986). Foams maintained their viscosity better under high temperatures than their base gel uids. The high-temperature stability
of foams depended more on surfactant type and concentration
rather than polymer concentration (Harris and Reidenbach, 1987).
Finer texture can be generated under higher shear rates, higher
surfactant concentrations, and higher pressures (Harris, 1989). Gas
types (N2 and CO2) only affected foam stability, but not rheology
(Harris, 1995). The foam viscosity can be increased by factors of
3 to 10 by cross-linking the polymers in the liquid phase of foams
(Harris, 1996).
Theoretical models have been developed for rheology of solid
suspensions, emulsions and foams. In the low quality regime, the
simplest model is a linear model that was proposed by Einstein
(Einstein, 1906) for dilute suspensions assuming unimodal size
spheres and ignoring particle interactions. The ratio of the viscosity of the suspensions (emulsion or foams) to the viscosity of the
continuous liquid phase, H(Q) is estimated to be

H (Q ) = 1 + Q + o (Q2)

Renolds Number, no unit


single phase volume, m3
foam quality, no unit
tube diameter, m

(1)

where Q is the quality (gas volume fraction), and is the Taylor


coefcient (Taylor, 1932). takes into account the impact of the
particle shape and the viscosity ratio between the internal and
external phases. For the same particle volume fraction, the higher
the coefcient, the larger is the suspension viscosity.

internal phase
external phase
gas phase of foam
liquid phase of foam
value at the wall
intrinsic value at the wall
apparent value at the wall
apparent value

=C

0.4 + in / ex
1 + in / ex

(2)

where min and mex are internal and external phase viscosity, respectively. The shape factor C is 2.5 for spheres and larger for nonspherical suspensions. For low quality foams, bubbles are spherical
and sparsely distributed in the uid, and the viscosity ratio of the
gas phase to the liquid phase is close to zero. So is close to 1. The
linear model is only valid for qualities lower than 10%. For higher
qualities, the most widely used models are those proposed by
Mooney (1951),

H (Q ) = exp (

Q
)
1 Q /Q m

(3)

and Quemada (1977),

H (Q ) = (1

Q Q m
)
Qm

(4)

where Qm is the critical quality. The critical quality is found experimentally to be the random close pack limit (Lee, 1970) which
is about 0.61 for spheres. Brouwers (2010) has proposed an
analytical expression for suspension viscosity with an unimodal
drop size,

H (Q ) = (

1Q
)Q m/ (1 Q m )
1 Q /Q m

(5)

As the quality increases, the bubbles touch each other to form


polyhedral structures with thin separating lamella. Within this
quality regime, the rheology behavior is controlled by the foam
structure and lamella properties. Princen and Kiss (1989) have
proposed an empirical model to correlate the volume fraction,
continuous phase viscosity, sauter mean diameter (SMD) and interfacial tension to the emulsion rheology, i.e.,

H (Q ) =

y
ex

+ 32 (Q 0.73) Ca1/2

(6)

where y is the yield stress, mex is the viscosity of the external


phase, and is the shear rate. Ca is the capillary number given as,

Ca =

ex d sv
2

(7)

where dsv is the SMD and s is the interfacial tension.


Herzhaft et al. (2005) measured the rheology for polymercontaining foam in a recirculating loop and compared the experimental results with several theories. Below a quality of 60%,

M. Gu, K.K. Mohanty / Journal of Petroleum Science and Engineering 134 (2015) 8796

the foam is Newtonian and the TaylorMooney (Mooney, 1951;


Taylor, 1932) viscosity model for moderately concentrated emulsions matches experimental results. For high volume fraction
foams ( Z 60%), the bubbles touch each other. The viscosity/shear
stress curves of these foams show a severe shear-thinning behavior beyond an apparent dynamic yield stress. The ow behavior
of high quality foams can be predicted by the Princen and Kiss
(1989) model, which incorporates bubble size distribution and
interfacial area.
Most of the studies above focused on the conventional fracturing foams with polymer additives as foam stabilizers. In our
study, to avoid the polymer damage in ultra-low permeable shales,
polymer-free aqueous foams are investigated. The goal of the
current work is to study the rheology of polymer-free foams under
a typical fracturing shear rate range (1001000 s  1). Foam ow
pressure drops are measured in a re-circulating loop and the effects of shear rate, foam quality, temperature, pressure, and chemical composition of the base uid are investigated. The experimental data has been compared with the theoretical models.
Empirical correlations are developed for the rheology of polymerfree foams, which can be incorporated in a fracture propagation
simulator to predict the fracture geometry and proppant distribution within the fractures.

89

Fig. 1. The recirculating loop for foam rheology measurement.

rotation speed and the corresponding pressure drops across the


test section are recorded after the steady state is achieved. All the
steps are repeated at several foam qualities. By conducting the
same test with different pipeline diameters, the wall slippage effect can be estimated. There is negligible slippage error in current
test.

2. Methodology
2.1. Fluids
The surfactant (AS) used in this paper is Bioterge AS40; it is an
anionic surfactant, 39% active as supplied. The viscoelastic surfactant (VES) was supplied by BJ Services. Glycerol (99% purity)
was used as a stabilizer in some of the tests. Four foam base uids
were formulated. Fluid A: 0.5 wt% anionic surfactant in water,
Fluid B: Fluid A 2 wt% glycerol, Fluid C: 0.5 wt% viscoelastic surfactant in water, Fluid D: 0.1 wt% anionic surfactant in water (a
diluted version of uid A). Nitrogen was used as the gas phase.
2.2. Foam loop
A circulating loop was built to generate foam at several quality,
temperatures and pressures. Foam was circulated in the loop at
several ow rates and corresponding pressure drops along a test
section (diameter 1.27 cm, length 15.24 m) were measured by a
differential pressure transducer. The whole loop was covered with
thermal insulators, except one part that was coated with a heating
jacket. The heating jacket can heat the circulating uid to a stable
temperature around 155 F. The pressure within the loop was set
by a back pressure regulator (PR) (connected to a gas cylinder) up
to a pressure of 2000 psi. A transparent view cell was included to
visually observe the foam texture. Fig 1 shows the schematic
diagram of the circulating loop.
The foam rheology measurement consists of the following
steps. The loop is lled with the surfactant solution by a gear
pump. Then, the system is heated to the desired temperature. A
certain back pressure is applied by introducing the N2 gas into the
loop. The drainage valve is opened and the liquid phase is drained
gradually through the back pressure regulator to the disposal tank
until a desired quality is achieved. The quality, which is also
known as gas volume fraction, is calculated from the mass balance
equation (l  mix)/(l  g). In the equation, the liquid density (l)
and gas density (g) are known, and the mixture density (mix) is
read from the densitometer integrated inside the mass ow meter.
Then the two-phase uid mixture is circulated at about 1000 s  1
shear rate to obtain homogeneous foam. It usually takes no more
than 10 min. Then the ow rates are varied by adjusting the pump

2.3. Rheology determination


Foam quality is dened as the percentage of gas volume in the
total volume of the foam at the in situ pressure and temperature,
i.e.,

Q=

100 Vg
Vg + Vl

(8)

The wall shear stress is calculated from the pressure drop (P)
as

w =

dP
4L

(9)

and the apparent wall shear rate is calculated from the ow


rate (q d2/4) as

wa =

8v
d

(10)

where d and L are the inner diameter and the length of the test
section. For a power law uid, the shear stress, is related to the
shear rate, by

= K n

(11)

where K is the consistency index and n is the power law index. In


Eq. (11), if shear stress is the wall shear stress w, the shear rate
should be the intrinsic shear rate at the wall wi, which can be
calculated as

wi =

(3n + 1) wa
4n

(12)

where n is given by the slope of the loglog plot of wall shear


stress w and the apparent wall shear rate wa, i.e.,

n =

d log (w )
d log (wa )

(13)

The power law index, n is equal to the parameter, n. The


consistency index, K can be calculated as

90

M. Gu, K.K. Mohanty / Journal of Petroleum Science and Engineering 134 (2015) 8796

(3n + 1) n
K = K /
4n

(14)

where K is the wall shear stress w at wa 1 s  1. The apparent


viscosity of the uid at any shear rate is given as

a =

w
= Kwi n 1
wi

(15)

The equations above are established under the conditions of


incompressible uid, fully developed laminar ow, and no slip
boundary.
HerschelBuckley model,

= 0 + K n

(16)

is used sometimes to describe foam rheology, but at a high


enough shear rate, the rst term on the right hand side is negligible and it is equivalent to a power law model (Eq. (11)). The base
uids A and B are Newtonian and their rheology is described by,

w = 8v2 Re1 for Re =

dv
2100

Fig. 2. Wall shear stress vs. apparent shear rate for foam A at 95 F.

(17)

in the laminar regime and

w = 0.0395v2 Re0.25 for Re =

dv
> 2100

(18)

in the turbulent regime. From Eqs. (10), (17) and (18), for laminar Newtonian ow

w wa

(19)

and for turbulent Newtonian ow

w wa1.75

(20)

because the exponent in Eq. (20) is greater than 1, a Newtonian


uid appears to be a shear thickening uid due to turbulence. The
base uid C has a viscoelastic surfactant, which forms entangled
chains under low shear rate and becomes aligned chains at high
shear rates. So it is shear thinning and follows Eqs. (11) or (16)
with n less than 1.

3. Results and discussions


Foam ow experiment was conducted with the three surfactant
formulations at varied qualities, temperatures, pressures, and
shear rates. Table 1 shows the range of test parameters. There
were four surfactant formulations: A, B, C, and D. Temperature was
varied from 95 to 155 F. The pressure was changed from 100 to
2000 psi. Foam quality was varied from 0% to 80%. The apparent
wall shear rate was varied from 100 to 1100 s  1. For each case,
ow rate and pressure drop were measured from which the wall
shear stress and apparent shear rate were calculated by Eqs.
(9) and (10), respectively. Fig. 2 shows wall shear stress versus
apparent shear rate for foams A for different qualities at 95 F. The
shear stress increases with increasing shear rate at the same

Fig. 3. Wall shear stress vs. apparent shear rate for foam B at 95 F.

quality. The shear stress increases with increasing quality at the


same shear rate. As the quality increases, the slope of the shear
stress-shear rate curve decreases, in general. Qualities up to 80%
were obtained for the uid A and B. The data for uid B, which is
shown in Fig. 3, is qualitatively similar to that of uid A. The slope
of the shear stress-shear rate curve is about 1.75 indicating that
the ow is turbulent for the liquid ow (quality 0%). At qualities
below 50%, at higher shear rate range, the ow is also turbulent.
For foam qualities between 50% and 60%, the slope is close to 1,
which shows the foam behaved like a Newtonian uid and the
ow is laminar in the whole shear rate range. But for foam qualities above 60%, the slope appears to be less than 1 indicating
shear thinning non-Newtonian ow behavior.
For the uid C, 060% qualities were successfully generated.
Above 63% quality, the foam C became too viscous to be mixed and
pumped in our ow loop. Fig. 4 shows wall shear stress versus
apparent shear rate for uid C foams for different qualities at 95 F.
The shear stress-shear rate curve for the pure liquid has a slope

Table 1
Test matrix.
No.

Liquid phase

Temp (F)

Pressure (psi)

Quality (%)

Pipe size (inch)

Mixing shear rate (s  1)

A
B
C
D

0.5 wt% AS (39%a)


0.5 wt% AS (39%a) 2 wt% Stabilizer (99%a)
0.5 wt% VES (99%a)
0.1 wt% AS (39%a)

95,125,155
95,125,155
95,125,155
95

1000
1000
1000
1000

080
00
080
080

0.5
0.5
0.5
0.5

1000
1000
1000
1000

Surfactant activity.

M. Gu, K.K. Mohanty / Journal of Petroleum Science and Engineering 134 (2015) 8796

91

Fig. 6. Apparent viscosity as a function of intrinsic shear rate for foam B at 95 F.


Fig. 4. Wall shear stress vs. apparent shear rate for foam C at 95 F.

less than 1, indicating shear thinning non-Newtonian behavior of


the base uid. The VES uid has a surfactant which aggregates into
cylindrical micelles which exhibits high and non-Newtonian
viscosity.
3.1. Effect of shear rates
The wall shear stress-apparent shear rate data were converted
to apparent viscosity versus intrinsic shear rate plots using Eqs.
(1215). Figs. 57 show the plots of apparent viscosity vs. intrinsic
shear rate for foam A, B and C at qualities of 080% and temperature of 95 F. For foams A and B, the viscosity decreases with
increasing shear rate for qualities above 60%, which is known as
the shear thinning behavior. This decrease is larger for higher foam
qualities. For low qualities ( o50%), apparent viscosity increases
slightly with the increase in shear rate. This increase in apparent
viscosity is due to the turbulent ow at the shear rates above
300400 s  1. The turbulence increases the ow resistance and
makes the foams behave like a shear thickening uid. The low
quality foams are dispersions of gas in water without much interaction between gas bubbles and the water ow is turbulent at
high shear rates. The pattern of behavior is similar for foams A and
B, because the base uid is Newtonian for these two cases. Unlike
foams A and B, foam C behaves like a shear-thinning uid from
low quality to high because of the shear-thinning nature of its liquid phase.
It is found that the low quality foams (2030%) generated from
uid A and B possess higher apparent viscosities at shear rate

Fig. 7. Apparent viscosity as a function of intrinsic shear rate for foam C at 95 F.

below 500 s  1 and lower apparent viscosities at shear rate above


500 s  1, compared with their base uids, as shown in Figs. 56.
That is because the bubbles dispersed in the base uid retard the
development of the turbulence at the high shear rate. This is a
good property for the fracturing uid, because during fracturing a
low friction loss in the tubing facility (high shear rate zone) and a
high proppant suspension viscosity in the fractures (low shear rate
zone) are always desired.
3.2. Effect of foam quality
Besides the shear rate effect, the effect of increasing foam
quality on the rheology of aqueous foam can also be investigated
from Figs. 57. The viscosity increases sharply with quality at the
high quality regime. For example, for foam A at a low shear rate
(100 s  1), the viscosity increases from 1.5 cp to 8 cp at the low
quality range (050%) and from 8 cp to 85 cp at the high quality
range (5080%). At a higher shear rate (1000 s  1), the viscosity
changes from 6 cp to 9.5 cp at the low quality range and 9.5 cp to
45 cp at the high quality range. This sudden increase of the viscosity above 50% quality is attributed to the transition from a
loosely packed bubble regime to a closely packed bubble regime as
quality increases.
3.3. Effect of temperature

Fig. 5. Apparent viscosity as a function of intrinsic shear rate for foam A at 95 F.

Figs. 810 show the effect of temperature on foam apparent


viscosities at 500 s  1 shear rate for uids A, B and C, respectively.
Generally the viscosity decreases with the increasing temperature.
According to the gures, the temperature effects are more

92

M. Gu, K.K. Mohanty / Journal of Petroleum Science and Engineering 134 (2015) 8796

Fig. 8. Effect of temperature on apparent viscosity for foam A at 500 s  1.


Fig. 11. Apparent viscosity vs. shear rate for base uids B and C.

Fig. 9. Effect of temperature on apparent viscosity for foam B at 500 s  1.


Fig. 12. Pressure dependence of the apparent viscosity of foam A at 500 s  1.

high temperatures. The apparent viscosity for base uid B hardly


changes as the temperature increases. Thus, increasing the temperature destabilizes the foam C more than other foams A and B by
thinning VES uid more than the regular surfactant uids.
3.4. Effect of pressure

Fig. 10. Effect of temperature on apparent viscosity for foam C at 500 s  1.

signicant for high quality foams than for low quality foams. Increasing temperature causes deterioration of the foams by accelerating the liquid drainage in the lamellae; in some cases the
surfactant solubility in the base solution can also decrease. The
viscosity of the VES foams (C) decreases more than the viscosity of
the regular surfactant foams (A and B) at high temperature. Fig. 11
shows the apparent viscosity versus shear rate for base uids B
and C with temperature. The viscosity of the VES uid
(C) decreases with increasing temperature for a given shear rate,
which is due to the degradation of the wormlike micelles of VES at

The effect of the pressure on the foam rheology is shown in


Fig. 12 for foam A at 500 s  1 shear rate and room temperature. For
low quality foam (o50%), the effect of pressure is small, but for
higher quality as the pressure increases the apparent viscosity
increases. The power law index, n and consistency index, K for
foam A at different pressures are shown in the Table 2. For the
foams with a quality below 50%, K and n change only slightly as
the pressure increases. This change is within the measurement
error. At higher quality, K and n are affected by the pressure. Fig. 13
shows that as pressure increases, the power law index, n decreases, but the rate of change decreases. Fig. 14 shows that as
pressure increases, the consistency index, K increases linearly.
Pressure affects the foam rheology by increasing gas densities,
slightly increasing gas viscosity, and most importantly changing
the foam textures (size distribution and mean diameter) (Harris,
1989; Herzhaft et al., 2005). Pressure has negligible effect on foam
rheology at low qualities because droplets have little effect on
rheology. Pressure has a signicant effect at high qualities because
the foam texture controls foam rheology.

M. Gu, K.K. Mohanty / Journal of Petroleum Science and Engineering 134 (2015) 8796

93

Table 2
Power law parameters for foam A at 95 F.
Quality
45%

59%

66%

74%

80%

n
K (Nsn/m2)
apn (cp)
n
K (Nsn/m2)
apn (cp)
n
K (Nsn/m2)
apn (cp)
n
K (Nsn/m2)
apn (cp)
n
K (Nsn/m2)
apn (cp)

200 psi

400 psi

600 psi

800 psi

1000 psi

1400 psi

1700 psi

1.16
2.72e  3
7.34
1.14
4.69e  3
10.93
1.07
9.12e  3
13.69
0.91
3.28e  2
18.11
0.85
1.14e  2
22.55

1.24
1.71e  3
7.49

1.24
1.74e  3
7.80
1.03
1.15e  2
13.94
0.89
4.10e  2
20.76
0.71
1.88e  1
31.48
0.59
4.75e  1
37.32

1.25
1.59e  3
7.52

1.24
1.86e  3
7.99
0.98
1.86e  2
15.85
0.84
6.35e  2
23.10
0.65
3.74e  1
41.60
0.49
1.19
50.19

0.91
2.85e  2
16.64
0.80
8.32e  2
24.09
0.62
5.07e 1
46.56
0.45
1.77
57.82

0.89
3.34e  2
17.08

Apparent viscosity is calculated under a typical fracturing shear rate of 500 s  1.

Fig. 13. Effect of pressure on the power law index n (solid dots: experiment data
from Table 2, lines: regression results, blank dots: predicted n from Eqs. (22) and
(23).

Fig. 15. Effect of foaming agents on apparent viscosity of foams at 95 F and


200 s  1.

concentration increases the foam viscosity, especially at high


qualities. The increased surfactant concentration not only enhances the foam lamella stability, but also increases the total interfacial area of the foam structure. Adding glycerol also increases
the foam viscosity by thickening the liquid phase. The polymerfree foams A and B are both less viscous than 0.24 wt% (20 lb per
1000 gallon) guar foams, while the VES foam C has an apparent
viscosity similar to that of guar foams (between 0.24 and
0.36 wt%). The formation of the wormlike micelles and their networks increase the viscosity of the liquid phase and the foam of
uid C signicantly.
3.6. Comparison with theoretical models
Fig. 14. Effect of pressure on the consistency index K (solid dots: experiment data
from Table 2, lines: regression results, blank dots: predicted K from Eqs. (22) and
(23).

3.5. Effect of the composition


Fig. 15 shows the apparent viscosity of foams produced from
several base uids and they were compared with the viscosity of
two conventional Guar foams from the literature (Sudhakar, 2003).
The condition is the same, with a pressure of 1000 psi, a temperature of 95 F, and a shear rate of 200 s  1. Comparing the results of foam A and foam D, increasing the surfactant

The experimental data are compared with the theoretical


models (Eqs. (16)) for the two foam quality regimes: a low foam
quality regime ( o60%) and a high foam quality regime ( Z60%).
For low quality foams, bubbles are spherical as shown in Fig. 16a
and sparsely distributed in the uid. Because the viscosity ratio of
the gas phase to the liquid phase is close to zero, is close to 1 in
Eq. (2). Fig. 17 shows the experimental data of the viscosity ratio, H
(Q) in comparison with the theoretical models. All the models and
the experimental data lie together in the quality range of 020%.
As the quality further increases, the analytical model proposed by
Brouwers (2010) agrees the best with the measured data, while
the Quemada (1977) model is the second best. The linear model

94

M. Gu, K.K. Mohanty / Journal of Petroleum Science and Engineering 134 (2015) 8796

Fig. 16. Foam texture images (Foam A, 95 F, 1000 psi).

Fig. 17. Viscosity ratio, H as function of foam quality Q as predicted by different


theoretical models and as measured (Foam A, 95 F, 1000 psi).

underestimates and the Mooney (1951) model overestimates.


As the quality increases, the bubble size increases and the lamella thickness decreases (Fig. 16b). Once they start to touch each
other, polyhedral structures are formed with separating ultra-thin
liquid layers (Fig. 16c). Within this quality regime, the rheology
behavior is controlled by the foam textures and lamella properties.
Assuming that foams and emulsions are similar, the Eqs. (6) and
(7) can, therefore, be applied to estimate the apparent foam viscosity, by interfacial tension with surface tension. Furthermore,
because the shear rate range in our experiments is above 100 s  1,
the rst term of Eq. (6) is negligible.
Fig. 18 plots the bubble size distribution of Foam A for different
foam qualities. The higher the foam quality, the broader is the size
distribution and the larger is the mean bubble size, consistent with
the measurements of Herzhaft et al. (2005). The SMD (sauter mean
diameter) is computed from the size distribution and plotted
against the foam quality in the smaller box within Fig. 18. The SMD
(in mm) can be expressed as,

SMD = 2.28Q2 3.11Q + 1.09

(21)

Fig. 19. Comparison of the experimental data and prediction by Princen and Kiss
model (Foam A, 95 F, 1000 psi).

for Q460%. The surface tension is 34 mN/m for 0.5 wt% and
38 mN/m for 0.1 wt% anionic surfactant uid, respectively. Fig. 19
compares the apparent viscosity predicted by Princen and Kiss
model (Eq. (6)) with the measured values. The results show that
the model has a good prediction for a 74% quality foam, while
overestimates the viscosity for an 83% quality foam by 3 times. The
discrepancy at higher qualities may be attributed to the broader
drop size distribution, which deviates from the single-size assumption in the model.

3.7. Correlation development


Having investigated the parameters that affect foam rheology,
correlations are developed to estimate foam rheology under typical eld conditions (shear rate, temperature and pressure) during fracturing. The power-law model parameters (n and k) are
obtained for each experiment by tting the experimental apparent
shear stress and rate; the results are presented in Table 3 along
with the correlation coefcient, R. Based on this table, correlations
are developed to estimate power-law (n) and consistency (k) indices for polymer-free foams.
Figs. 20 and 21 show the power-law index (n) and the consistency index (k) as functions of quality for foam A at 1000 psi
pressure and different temperatures. From the plots, it is observed
that the rheology of the surfactant foam A depends on quality, but
is not very sensitive to the temperature. The power-law index
decreases with an increase in quality, while the consistency index
increases exponentially with the foam quality. These variations
can be represented by the empirical correlations listed below
(dashed lines in Figs. 20 and 21):

n0 (Q ) = 1.54 1.64Q2
Fig. 18. Bubble size distribution (Foam A, 95 F, 1000 psi).

(22)

M. Gu, K.K. Mohanty / Journal of Petroleum Science and Engineering 134 (2015) 8796

95

Table 3
Power law model parameters for foams A, B and C at 1000 psi.
T (F)

Parameter

95

125

155

95

125

155

95

125

155

Quality (%)
0

2030

3040

5060

6070

7075

7580

8085

n
K (Nsn/m2)
R2
n
K(Nsn/m2)
R2
n
K (Nsn/m2)
R2

1.60
1.00e  4
0.998
1.54
1.30e  4
0.995
1.52
1.40e  4
0.997

1.25
9.10e  4
0.996
1.35
4.50e  4
0.992
1.39
3.20e  4
0.992

1.07
5.75e  3
0.994
1.12
3.63e  3
0.994
1.16
2.13e  3
0.996

0.88
4.26e  2
0.999
0.89
3.55e  2
0.999
0.91
2.58e  2
0.999

0.64
3.98e  1
0.999
0.54
7.08e  1
0.996
0.51
8.58e  1
0.999

0.53
1.15
0.998
0.45
1.68
0.997
0.42
1.94
0.998

n
K (Nsn/m2)
R2
n
K (Nsn/m2)
R2
n
K (Nsn/m2)
R2

1.63
8.00e  5
0.9990
1.56
1.30e  4
0.997
1.53
1.50e  4
0.998

1.34
4.90e  4
0.99255
1.43
2.6e  4
0.994
1.51
1.50e  4
0.991

1.14
4.03e  3
0.998
1.24
1.79e  3
0.999
1.23
1.36e  3
0.999

0.79
9.64e  2
0.999
0.78
8.56e  2
0.998
0.78
8.02e  2
0.998

0.61
5.45e  1
0.997
0.46
1.47
0.992
0.45
1.45
0.999

0.57
9.32e  1
0.999
0.45
1.70
0.999
0.41
2.03
0.994

0.55
1.36
0.997
0.44
2.27
0.995
0.40
2.76
0.998

n
K (Nsn/m2)
R2
n
K (Nsn/m2)
R2
n
K (Nsn/m2)
R2

0.52
0.19
0.992
0.75
2.98e  2
0.997
0.92
8.24e 3
0.999

0.47
0.43
0.991
0.62
0.12
0.994
0.91
1.66e  2
0.999

0.35
1.79
0.997
0.51
0.62
0.988
0.74
7.91e  2
0.999

0.28
4.21
0.999
0.49
0.98
0.993
0.63
0.25
0.992

Fitting Corr.:K=10^(5.89Q2+0.43Q-4)

Fig. 20. Power law index n for foam A.


Fig. 21. Consistency index K for foam A.
2
K0 (Q ) = 10(5.89Q + 0.43Q 4)

(23)

These correlations can be used to estimate the foam rheology


for quality less than 60%. For higher quality, the inuence of
pressure needs to be considered. The effect of pressure on n and K
can be captured by the following correlations:

n (p, Q ) = n0 (Q ) + (0.21 0.89Q ) [log (P /1000)]

K (p, Q ) = K0 (Q ) + 8.6 1011e21Q [P 1000]

(24)

(25)

where P is the pressure, and n0, K0, are power law parameters
predicted by Eqs. (22) and (23). The correlations are shown as lines
in Figs. 13 and 14. The foam quality range normally used in

conventional reservoirs is 7095%. Lower quality foam can be


used in shales.
Our test equipment is limited to temperatures below 150 F and
pressures below 2000 psi, which is relevant to shallow formations.
The correlations are still good approximations at higher pressures
for foam quality lower than 60% because pressure has negligible
impact on foam rheology based on the previous discussion. For
foam quality higher than 60%, one can extrapolate the correlations
for n and K to higher pressures. The extrapolations should be validated by lab experiments at high pressures.
4. Conclusions
In this paper, the rheology of three kinds of polymer-free foams

96

M. Gu, K.K. Mohanty / Journal of Petroleum Science and Engineering 134 (2015) 8796

(A: 0.5 wt% regular anionic surfactant, B: Fluid A 2 wt% glycerol,


C: 0.5 wt% viscoelastic surfactant) are studied in a recirculating
pipe rheometer at temperatures below 150 F and pressures below
2000 psi.

 All three foams exhibit power-law rheological behavior. The




regular surfactant foams (A and B) show shear thinning behavior at qualities above 60%, non-shear dependent behavior from
50% to 60% and shear thickening below 50% (due to turbulence).
The VES foams show shear thinning at qualities less than 60%.
Temperature lowers the viscosity of foams due to decrease in
the liquid viscosity and increasing instability of the bubbles, but
the temperature effect is small for foams A and B. Pressure increases foam viscosity; the increase is higher for higher quality
foams. As the pressure increases, the rate of viscosity increase
decreases.
The aqueous foams A and B are both less viscous than 0.24 wt%
polymer (guar) foams, while the VES foam C has an apparent
viscosity similar to that of 0.240.36 wt% guar foams.
The model proposed by Brouwers (2010) agrees the best with
the measured data for quality under 60%; the model proposed
by Princen and Kiss (1989) for high quality does not match the
experimental data.
New correlations have been developed to describe the aqueous
foam rheology as a function of shear rate, quality, and pressure
at the parameter range typical of hydraulic fracturing. The correlations can be incorporated in a fracture propagation simulator to evaluate the foam fracturing efciency.

Acknowledgments
We thank Statoil and RPSEA (Grant No. 07122-38) for partial
funding of this work and BJ Services for the supply of the viscoelastic surfactant.

References
Bonilla, L.F., Shah, S.N., 2000. Experimental investigation on the rheology of foams.
SPE 59752. In: Proceedings of SPE/CERI Gas Technology Symposium. 35 April
2000. Calgary, Canada.
Brannon, H.D., Starks, T.R., 2009. Maximizing return-on-fracturing-investment by
using ultra-lightweight proppants to optimize effective fracture area: can less

be more? SPE 119385. In: Proceedings of SPE Hydraulic Fracturing Technology


Conference. 1921 January 2009. The Woodlands, Texas.
Brouwers, H.J.H., 2010. Viscosity of a concentrated suspension of rigid monosized
particles. Phys. Rev. E 81, 051402.
Cipolla, C.L., Warpinski, N.R., Mayerhofer, M.J., Lolon, E.P., Vincent, M.C., 2010. The
relationship between fracture complexity, reservoir properties, and fracturetreatment design. SPE Prod. Op. 25 (4), 438452.
Einstein, A., 1906. Effect of suspended rigid spheres on viscosity. Ann. Phys. 19,
289306.
Gidley, J.L., Holditch, S.A., Nierode, D.E., Veatch Jr., R.W., 1990. Recent Advances in
Hydraulic Fracturing. SPE Monograph 12. Society of Petroleum Engineers, Richardson, TX.
Harris, P.C., Reidenbach, V.G., 1987. High-temperature rheological study of foam
fracturing uids. J. Pet. Technol. 39 (5), 613619.
Harris, P.C., 1989. Effects of texture on rheology of foam fracturing uids. SPE Prod.
Eng. 4 (3), 249257.
Harris, P.C., 1995. A comparison of mixed gas foams with N2 and CO2 foam fracturing uids on a ow loop viscometer. SPE Prod. Facil. 10 (3), 197203.
Harris, P.C., 1996. Rheology of crosslinked foams. SPE Prod. Facil. 11 (2), 113116.
Herzhaft, B., Kakadjian, S., Moan, M., 2005. Measurement and modeling of the ow
behavior of aqueous foams using a recirculating pipe rheometer. Colloids Surf.
A: Physicochem. Eng. Asp. 263, 153164.
Hutchins, R.D., 2005. A circulating-foam loop for evaluating foam at conditions of
use. SPE Prod. Facil. 20 (4), 286294.
King, G.E., 2010. Thirty years of gas shale fracturing: what have we learned? SPE
133456. In: Proceedings of SPE Annual Technical Conference and Exhibition.
1922 September 2010. Florence, Italy.
Lee, D.I., 1970. Packing of spheres and its effect on the viscosity of suspensions. J.
Paint Technol. 42, 579587.
Mooney, M., 1951. The viscosity of a concentrated suspension of spherical particles.
J. Colloid Sci. 6, 162170.
Peles, J., Wardlow, R.W., Cox, G., Haley, W., Dusterhoft, R., Walters, H.G., Weaver, J.,
2002. Maximizing well production with unique low molecular weight frac uid.
SPE 77746. In: Proceedings of SPE Annual Technical Conference and Exhibition.
29th September to 2nd October 2002. San Antonio, Texas.
Princen, H.M., Kiss, A.D., 1989. Rheology of foams and concentrated emulsions an
experiment study of the shear viscosity and yield stress of concentrated
emulsions. J. Colloid Interface Sci. 128, 176186.
Quemada, D., 1977. Rheology of concentrated disperse systems and minimum energy dissipation principle. Rheol. Acta 16, 8294.
Reidenbach, V.G., Harris, P.C., Lee, Y.N., Lord, D.L., 1986. Rheological study of foam
fracturing uids using nitrogen and carbon dioxide. SPE Prod. Eng. 1 (1), 3141.
Sani, A.M., Shah, S.N., Baldwin, L., 2001. Experimental investigation of xanthan foam
rheology. SPE 67263. In: Proceedings of SPE Production and Operations Symposium. 2427 March 2001. Oklahoma City, Oklahoma.
Sudhakar, D.K., 2002. New empirical friction loss correlation for foam uids in
coiled tubing. SPE 74810. In: Proceedings of SPE/ICoTA Coiled Tubing Conference and Exhibition. 910 April 2002. Houston, Texas.
Sudhakar, D.K., 2003. New rheological correlations for guar foam uids. SPE 88032.
In: Proceedings of SPE Production and Operations Symposium. 2225 March
2003. Oklahoma City, Oklahoma.
Taylor, G.I., 1932. The viscosity of a uid containing small drops of another uid.
Proc. R. Soc. Lond Ser. A 138, 4145.
Warpinski, N.R., Mayerhofer, M.J., Vincent, M.C., Cipolla, C.L., Lolon, E.P., 2009. Stimulating unconventional reservoir: maximizing network growth while optimizing fracture conductivity. J. Can. Pet. Technol. 48 (10), 3951.

Potrebbero piacerti anche