Analytical Analysis of Fracture Conductivity For Sparse Distribution of Proppant Packs

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of Geophysics and Engineering

PAPER Related content


- Transient pressure behavior for a
Analytical analysis of fracture conductivity for horizontal well with multiple finite-
conductivity fractures in tight reservoirs
sparse distribution of proppant packs Jingjing Guo, Haitao Wang and Liehui
Zhang

- Experimental and numerical investigations


To cite this article: Jianchun Guo et al 2017 J. Geophys. Eng. 14 599 on the vertical propagation of hydraulic
fractures in laminated shales
Qinglin Shan, Yan Jin, Peng Tan et al.

- Transient pressure analysis of a volume


fracturing well in fractured tight oil
View the article online for updates and enhancements. reservoirs
Cheng Lu, Jiahang Wang, Cong Zhang et
al.

Recent citations
- Crushing characteristics of four different
proppants and implications for fracture
conductivity
Wenbo Zheng et al

- Influence of gravel on the propagation


pattern of hydraulic fracture in the glutenite
reservoir
Zhenhua Rui et al

This content was downloaded from IP address 125.70.254.70 on 25/05/2018 at 04:14


Journal of Geophysics and Engineering

J. Geophys. Eng. 14 (2017) 599–610 (12pp) https://doi.org/10.1088/1742-2140/aa6215

Analytical analysis of fracture conductivity


for sparse distribution of proppant packs
Jianchun Guo1, Jiandong Wang1,2, Yuxuan Liu1, Zhangxin Chen3 and
Haiyan Zhu1
1
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation, Southwest Petroleum
University, Chengdu 610500, People’s Republic of China
2
Sinopec Shengli Oilfield Company, Shengli 257001, People’s Republic of China
3
Department of Chemical and Petroleum Engineering, Schulich School of Engineering, University of
Calgary, Calgary T2N 1N4, Canada

E-mail: guojianchun@vip.163.com

Received 28 March 2016, revised 13 December 2016


Accepted for publication 22 February 2017
Published 4 April 2017

Abstract
Conductivity optimization is important for hydraulic fracturing due to its key roles in
determining fractured well productivity. Proppant embedment is an important mechanism that
could cause a remarkable reduction in fracture width and, thus, damage the fracture
conductivity. In this work a new analytical model, based on contact mechanics and the
Carman–Kozeny model, is developed to calculate the embedment and conductivity for the
sparse distribution of proppant packs. Features and controlling factors of embedment, residual
width and conductivity are analyzed. The results indicate an optimum distance between
proppant packs that has the potential to maintain the maximum conductivity after proppant
embedment under a sparse distribution condition. A change in the optimum distance is
primarily controlled by closure pressure, the rock elastic modulus and the proppant elastic
modulus. The proppant concentrations and the poroelastic effect do not influence this
optimum distance.
Keywords: proppant, embedment, conductivity, hydraulic fracturing

(Some figures may appear in colour only in the online journal)

Nomenclature
Fl Force located at Lhf<x<Lhf, N m−1
2
Af Area of fracture face, m Fc Fracture conductivity, m2.m
a Radius of contact area, m Fi Force located at part i, N m-1
2
A Flow area, m f Fitting parameters, dimensionless
Af Area of rock face, m2 H Length in the direction of rows, m
As Contact area of single particle, m2 Hp Length of proppant pack in the direction of
C Kozeny–Carman constant, dimensionless rows, m
D2 Thickness of the upper and lower rocks, m h Embedment depth, m
ΔD1 Deformation of proppant packs, m J Layers of proppants, dimensionless
ΔD2 Deformation of rock, m K Stress intensity factor, Pa-m0.5
E1 Elastic modulus of proppant, Pa Ki Stress intensity factor at fracture tip induced
E2 Elastic modulus of rock, Pa by part i, Pa-m0.5
F Applied force on proppant, N k Permeability, m2

1742-2132/17/030599+12$33.00 599 © 2017 Sinopec Geophysical Research Institute Printed in the UK


J. Geophys. Eng. 14 (2017) 599 J Guo et al

L Length in the direction of columns, m


Lp Length of proppant pack in the direction of
columns, m
Lhf Fracture half length, m
M Columns of proppants, dimensionless
N Rows of proppants, dimensionless
NHF Number of parts divided in hydraulic fracture,
dimensionless
pmax Maximum contact pressure Pa
pave Average contact pressure, Pa
pc Closure pressure, Pa
po Overburden pressure, Pa Figure 1. Illustration of a hydraulic fracture.

pf Fluid pressure in the fracture, Pa


ps Pressure of single proppant acting on rock, Pa
Δp Pressure difference, Pa
Δptt Pressure drop from fracture tip to fracture
toe, Pa
q Flow rate, m3 s−1
R Radius of proppant, m
Rsi(i=1, 2) Proppant distance ratio, dimensionless
u Velocity, m s−1
wf Fracture width, m
wf0 Initial width of fracture, m
Δwf Change of fracture width, m Figure 2. Width reduction due to proppant embedment.
x Position along fracture length, m
v1 Poisson’s ratio of proppant, dimensionless
v2 Poisson’s ratio of rock, dimensionless studied (Clark and Guler 1983, Mobbs and Hammond 2001,
δ Indentation depth, m Dontsov and Peirce 2014). In unconventional reservoirs, slick
water is widely used to generate the fracture network. The
μ Viscosity of fluid, Pa.s
fluid has low viscosity and poor proppant carrying capacity,
f Porosity of the sample, dimensionless resulting in some nonuniform distributions of proppants in the
fracture network (Awoleke et al 2016). Experimental and
field tests both show that sparse distribution—such as a partial
1. Introduction monolayer in secondary fractures (Fredd et al 2001, Zou
et al 2015) or sparse distribution in channel fracturing
Hydraulic fracturing is a reservoir stimulation method often (D’Huteau et al 2011)—can improve the fracture con-
applied to low permeability formations that would otherwise ductivity. The variation of fracture conductivity under sparse
be uneconomical to produce. The process starts with the high- distribution, therefore, is very important in optimizing the
pressure injection of a fracturing fluid into a wellbore to fracturing design in unconventional reservoirs.
create cracks. Fluids containing sand or other proppants Proppant embedment may significantly impair fracture
suspended with the aid of thickening agents are then injected.
conductivity due to a decrease in the fracture width (figure 2).
When the hydraulic pressure is removed from a well, the sides
Proppant embedment and deformation play significant roles
of the fractures compress onto the proppants, creating a high-
in the stimulation response of hydraulic fracturing, especially
permeability pathway through which natural gas and petro-
leum flow more freely in deep-rock formations (figure 1). in clay-rich, soft formations. Additionally, a stress shadow
The response of fractured wells is highly dependent on effect, caused by multiple fracture interactions, can increase
fracture conductivity. In conventional reservoirs, a high the effective closure stress (Jin et al 2010), resulting in
viscosity crosslinked fluid is used to create long and sym- enhanced proppant embedment. This is generally regarded as
metric fractures. Proppants are believed to be placed uni- a geomechanical effect in an unconventional reservoir simu-
formly in the fracture due to high viscosity and good proppant lation and should be considered when assessing the produc-
carrying capacity (El-M. Shokir and Al-Quraishi 2009). tion forecast in formations with low Young’s modulus (Yu
Factors affecting the proppant placement have been widely and Sepehrnoori 2014).

600
J. Geophys. Eng. 14 (2017) 599 J Guo et al

Table 1. Factors considered in different models.

Model Rock deformation Proppant deformation Sparse distribution of conductivity Conductivity model type
Guo and Liu (2012) √ Analytical
Khanna et al (2012) √ √ Numerical
Neto et al (2015) √ √ Semi-analytical
Li et al (2014) √ √ Analytical
Model in this paper √ √ √ Analytical

Currently, it is still impossible to conduct in situ mea- 2. Model development


surements of residual width and conductivity. Several
experimental apparatus were developed to study the features In hydraulic fracture, proppants are packed in multilayers, as
of residual width and conductivity (Volk et al 1981, Hartley shown in figure 3. J layers of proppants support a fracture
and Bosma 1985, Huitt and McGlothlin 1986, Lacy width of wf. In the top view, the M columns and N rows of
et al 1997, Lacy et al 1998, Wen et al 2007, Guo et al 2008, proppants result in an area of the fracture region Af=L×H.
Zhang et al 2015). Based on the results of these experiments,
empirical formulas to calculate residual width and con- 2.1. Theoretical background
ductivity were proposed. Heinrich Hertz has derived the solution of the contact between
In recent years, several analytical models have also been an elastic sphere and an elastic half-space (Johnson 1987),
developed. Guo and Liu (2012) proposed an analytical model as illustrated in figure 4. An elastic sphere of radius R indents
to describe the embedment features in soft rock. Their model, an elastic half-space to depth δ, and thus creates a contact area
however, only considered closely packed cases. Khanna et al of radius
(2012) proposed a simplified approach for calculating the
conductivity of narrow fractures filled with a sparse mono- a= Rd . (1 )
layer of proppant particles. The particle is considered to be a
The applied force F is related to the displacement depth δ by
rigid body. Neto et al (2015) developed a semi-analytical
model for residual width and conductivity based on a dis- 4
F= E *R 2 d 2
1 3
(2 )
tributed dislocation technique with consideration of proppant 3
pack deformation, which is described by Terzaghi’s classical
consolidation model. The sparse distribution of proppant was where
not considered in this model. Li et al (2014) developed an 1 1 - v12 1 - v22
analytical model to calculate proppant embedment based on = + (3 )
E* E1 E2
contact mechanics for sparse placement of proppant. The
Carman–Kozeny model was modified to predict the con- where v1 is Poisson’s ratio of proppants (dimensionless); v2 is
ductivity for proppant packs, although sparse distribution was Poisson’s ratio of rock (dimensionless); E1 is the elastic
not considered. In summary, theoretical studies on proppant modulus of proppants (Pa); E2 is the elastic modulus of
embedment and conductivity have been widely conducted rock (Pa).
(table 1); however, conductivity under sparse distribution has The radius of the contact circle a, depth δ and the max-
received limited attention. The characteristics and controlling imum contact pressure pmax are given by
factors of conductivity under sparse distribution is not clear, 1
especially for multi-layer proppant packs. ⎛ 3FR ⎞ 3
a= ⎜ ⎟ (4 )
For this paper, a new analytical model, based on contact ⎝ 4E * ⎠
mechanics and the Carman–Kozeny model, was developed to
1
address this problem. A number of simplifications were ⎛ 9F 2 ⎞ 3
d=⎜ ⎟ (5 )
adopted, including disregarding the surface roughness of ⎝ 16RE *2 ⎠
fractures, possible proppant crush and secondary cracking of
fracture walls due to very high contact stresses. The proppant ⎛ 6FE *2 ⎞ 3
1

embedment and deformation are considered to be the primary pmax =⎜ 3 2 ⎟ . (6 )


⎝ pR ⎠
mechanism of fracture width change and fracture conductivity
loss. These assumptions limit the developed approach’s The average contact pressure compressing proppants is rela-
ability to accurately predict the conductivity. Even so, the ted to the maximum contact pressure
results can provide insight into the features and factors that
control the conductivity change induced by proppant 2
pave = pmax . (7 )
embedment and deformation. 3

601
J. Geophys. Eng. 14 (2017) 599 J Guo et al

Figure 3. Problem geometry.

As = pa2 . (11)

Combining equations (8)–(11), ps can be derived as

4pc R 2
ps = (12)
pa2

and

ps = pave . (13)

Combining equations (7), (12) and (13), the maximum


contact pressure pmax can be expressed as
Figure 4. Contact of an elastic sphere with an elastic half-space.
3 4pc R 2
pmax = (14)
2 pa2

Combining equations (4)–(6) and (14), the depth δ and


maximum contact pressure pmax can be derived as
2
⎛ 3p ⎞ 3
d = ⎜ c⎟ R (15)
⎝ E* ⎠

1 1
pmax = (24pc E *2) 3 . (16)
p

Figure 5. Stress analysis of proppants.


The thickness of the upper and lower rocks is the same
and assumed as D2. The deformation of rock is reflected in the
2.2. Even distribution of proppants form of proppant embedment (Li et al 2014). Therefore, the
2.2.1. Residual width. Figure 5 shows a stress analysis of a
embedment depth h can be expressed as
proppant pack. The relationships between the load and stress 2
satisfy the following equations: ⎛ 3p ⎞ 3 p
h = d + DD2 = ⎜ c ⎟ R + c D2 (17)
⎝ E* ⎠ E2
pc Af = ps MNAs (8 )
and where h is the embedment depth, (m); ΔD2 is the deformation
of rock (m).
pc = po - pf (9 )
The deformation of proppant packs satisfies the follow-
where pc is the closure pressure (Pa); po is the overburden ing equation:
pressure (Pa); pf is the fluid pressure in a fracture (Pa); ps is
pave
the pressure with which a single proppant acts on rock (Pa); DD1 = wf0 (18)
Af is the area of the propped fracture surface (m2); As is the E1
contact area of a single particle (m2). Af and As can be
where ΔD1 is the deformation of proppant packs (m); wf0 is
expressed as follows:
the initial width of a fracture (m). Combining equations (7),
Af = 4MNR 2 (10) (17) and (18), the change in the fracture width Δwf can be

602
J. Geophys. Eng. 14 (2017) 599 J Guo et al

Figure 6. Illustration of sparse distribution of proppant packs.

obtained by
2
2pmax ⎛ 3p ⎞ 3 Figure 7. A representative unit in figure 6.
Dwf = DD1 + 2h = wf0 + 2 ⎜ c ⎟ R
3E1 ⎝ E* ⎠
2pc
+ D2 (19) Using the same derivation process it is found that in even
E2
distribution, the depth δ and maximum contact pressure can
and be derived as
wf = wf0 - Dwf (20) 2
⎛ 3pc ⎞ 3
where Δwf is the change in fracture width (m); wf is the final d=⎜ ⎟ R (25)
⎝ E *Rs1Rs2 ⎠
fracture width (m).
1
2.2.2. Fracture conductivity. Results from discrete element 1 ⎛ 24pc E *2 ⎞ 3
pmax = ⎜ ⎟ . (26)
modeling indicate that the Carman–Kozeny model can be p ⎝ Rs1Rs2 ⎠
used to calculate the permeability of proppant packs
The embedment depth, h, and the change in fracture width,
(Mollanouri Shamsi et al 2015, Sanematsu et al 2015). The
Δwf, can be obtained as
permeability is described as follows:
f3 ⎛ 3pc ⎞ 3
2
kf = (2 R )2 (21) h = d + DD2 = ⎜
p
⎟ R + c D2 (27)
36C (1 - f)2 ⎝ E *Rs1Rs2 ⎠ E2
where f is the porosity of a sample; C is the Carman–Kozeny
2
constant. For beds packed with spherical particles, the constant 2p ⎛ 3pc ⎞ 3
C is equal to 5. The fracture conductivity can be obtained by Dwf = DD1 + 2h = max wf0 + ⎜ ⎟ R
3E1 ⎝ E *Rs1Rs2 ⎠
Fc = k f ´ wf . (22) p
+ c D2 . (28)
E2
2.3. Uneven distribution of proppants
2.3.2. Fracture conductivity. A representative unit in figure 6
2.3.1. Residual width. The proppant packs may have a sparse
is shown in figure 7. Conductivity of this unit is developed in
distribution in a hydraulic fracture, especially in channel
this section. The other unit can be derived using the same
fracturing (D’Huteau et al 2011). A simple sparse distribution
process. In this figure, q is the flow rate (m3 s−1), u is the
is illustrated in figure 6.
velocity (m s−1), k is the permeability (m2), and Δp is the
A two-proppant distance ratio (Rs1, Rs2) can be defined as
pressure difference (Pa). The subscripts (1, 2, 3) represent
Hp
2NR the different sections in this figure.
Rs1 = = (23)
H
H According to Darcy’s Law
Lp
2MR
Rs2 = = (24) kA dp
L L q = uA = - (29)
m dx
where Lp and Hp are the length and height of a proppant pack
(m). If the two packs are close to each other, Rs1 and Rs2 are where A is the flow area (m2); μ is the viscosity of a fluid
equal to unity; if not, they are smaller than unity. (Pa.s). According to the relationship between the pressure

603
J. Geophys. Eng. 14 (2017) 599 J Guo et al

Table 2. Basic parameters.

Data source v1 v2 E1 (MPa) E2 (MPa) R (mm) D2 (mm)


Lacy et al (1998) 0.13 0.13 21 306 1172 0.3175 20
Lu et al (2008) 0.144 0.144 35 000 28 770 0.3175 15

difference in sections 1 and 2, the equivalent permeability of


sections 1 and 2 is

1 1 1
1,2
= (1 - Rs2) + Rs2 . (30)
k eq k1 k2

According to the relationship between the flow rate in


sections 1 and 3, the equivalent permeability of the three
sections is
1,2,3
k eq = k eq
1,2
Rs1 + k3 (1 - Rs1) . (31)

The permeability of the different sections is


Figure 8. Comparison of theoretical values of proppant embedment
with experimental data from Lacy et al (1998).
f3
k2 = d2
36C (1 - f)2
wf2
k1 = k3 = . (32)
12

Subsequently, fracture conductivity can be obtained by

Fc = k eq
1,2,3
´ wf . (33)

3. Discussions

3.1. Model verification

To verify the new model, it is necessary to compare the new


Figure 9. Comparison of modified models and experimental data
model with the existing model and experimental data. Lacy from Lacy et al (1998).
et al (1998) and Lu et al (2008) measured the proppant
embedment and width change at different closure pressures. Table 3. Fitting results.
The experimental data they reported were used to test the new
Lacy et al (1998) Lu et al (2008)
model in this paper and the model developed by Li et al
2
(2014). Li et al (2014) summarized the known experimental Model f R f R2
conditions as listed in table 2. New model 1.314 0.92 16.49 0.97
The comparison of the theoretical values of proppant Li et al (2014) 1.322 0.92 21.14 0.95
embedment with experimental data from Lacy et al (1998) is
shown in figure 8. The values calculated with the new model
and the model by Li et al (2014) are very close. The two The same fitting parameters were introduced into the
theoretical models’ values are smaller than the experimental model developed by Li et al (2014). The comparisons of the
data. The reason for the smaller values has been discussed by modified models are shown in figure 9. The revised models fit
Li et al (2014). well with the experimental data and the calculated values of
The fitting parameters were introduced in equation (17): these two models almost overlap with each other. The cor-
rection factors and regression coefficients of the revised
⎡ 2 ⎤ models in figure 9 are listed in table 3.
⎛ 3p ⎞ 3 p
h = f ⎢ ⎜ c ⎟ R + c D2 ⎥ . (34) The same fitting parameters were introduced into the data
⎢⎣ ⎝ E * ⎠ E2 ⎥⎦
provided by Lu et al (2014). The comparisons of the modified

604
J. Geophys. Eng. 14 (2017) 599 J Guo et al

Figure 10. Comparisons of modified models and experimental data Figure 13. Embedment under different proppant distance ratios.
from Lu et al (2008).
The comparisons of the modified models and experimental
data from Lacy et al (1998) are shown in figure 11. The two
modified models’ values are both smaller than the exper-
imental data; the deviation of the model values from the
experimental data, however, is not large and follows the trend
of the experimental data.
The comparisons of the modified models and experimental
data from Lu et al (2008) are shown in figure 12. The new
modified model in this paper can match most of the data until
the closure pressure increases to 40 MPa. At the same time,
the modified model results of Li et al (2014) are smaller than
the experimental data. The results demonstrate that the new
modified model can be used to calculate the width change.

Figure 11. Comparisons of the modified models and experimental


data from Lacy et al (1998). 3.2. Application to the sparse distribution of proppant packs

3.2.1. Changes in embedment, width and conductivity. In this


section, the new models modified by two sets of experimental
data are used to predict the embedment, fracture width change
and conductivity in the sparse distribution of proppant packs.
The results of the new model modified by the data from Lacy
et al (1998) is referred to as case A; the other, with the data
from Lu et al (2008), is referred to as case B.
Embedment can be obtained with the modified new
model at closure pressure Pc=30 MPa, as shown in
figure 13. With an increase in Rs1, embedment decreases, as
expected. The change in embedment is greater when Rs1 is
small. When Rs1 reaches a specific value, the change in
embedment is very small. For values in case A, when Rs1
increases to 0.4, the embedment depth reduces by more than
Figure 12. Comparisons of the modified models and experimental 30%; when the value increases from 0.4 to 1, the embedment
data from Lu et al (2008). depth changes only 8%. This shows that with an increase in
the proppant distance ratio, the embedment depth, as well as
models are shown in figure 10. Similarly, the two models fit the reduction amplitude, decreases. For values in case B the
well with the experimental data; the fitting parameters are also embedment depthalmost reduces by half when Rs1 increases
listed in table 3. The results indicate that the new model can to 0.2; when Rs1 increases from 0.2 to 1, the embedment
match the experimental data in all the cases studied, with the depth changes by 42%.
same or slightly better fitting results as compared with the Figure 14 shows the relationship between a fracture
model by Li et al (2014). width change and Rs1 for the two cases. The trend is similar to
A fracture width change is required in the calculation of that of proppant embedment. With an increase in Rs1, the
fracture conductivity. Consequently, the width change is also fracture width decreases. Its change is greater when Rs1 is
compared with the modified model and experimental data. small because, with an increase in Rs1, the pressure applied on

605
J. Geophys. Eng. 14 (2017) 599 J Guo et al

Figure 14. Fracture width change under different proppant distance Figure 15. Relationship between stress intensity factor and Rs1. The
ratios. stress intensity factor increases with the increased proppant distance
ratio.
the rock and proppant packs reduces. From figures 13 and 14,
the trends indicate that Rs1 has a significant influence on
embedment and fracture width.
Furthermore, the stress intensity factor at the crack tip
would yield useful information on fracture closure especially
when the proppants are unevenly and sparsely distributed
within the fracture. The stress intensity factor (K ) is used in
fracture mechanics to predict the stress state near the fracture
tip. When the fracture is open, a positive value of K exists at
the fracture tip. Otherwise, a negative or zero value of K
indicates the fracture closure.
The stress intensity factor K at the fracture tip by a point
force F1 can be obtained from the following expression
(Wells 1965):
Fl L Hf LHf + x
ò-L
Figure 16. Fracture conductivity under different proppant distance
K= dx . (35)
2 pLHf Hf LHf - x ratios. There is an optimal distance for maximum conductivity after
proppant embedment and deformation.

An approximate formula can be expressed as obtained by equation (36) and then the integrated value can be
achieved by equation (37).
Fi LHf + x
Ki = (36) Figure 15 shows the relationship between the stress
2 pLHf LHf - x intensity factor and Rs1 at different fracture lengths. The Kn is
NHF a normalized stress intensity factor, which is normalized
K= å Ki (37) against K0. The relationships presented in figure 15 clearly
i=1 show the expected tendency of K approaching K0, when the
where K is the stress intensity factor (Pa-m0.5); Ki is the stress fracture is fully filled with proppants (Rs1=1). This case also
intensity factor at the fracture tip induced by part i (Pa-m0.5); corresponds to the maximum residual width of the fracture.
Fl is the force located at Lhf<x<Lhf (N m−1); Fi is the Fracture conductivity after proppant embedment and
force located at part i (N m−1); Lhf is the fracture half length deformation can also be obtained, as shown in figure 16. The
(m); x is the position along the fracture length (m); NHF is the fracture conductivity is normalized against the conductivity of
number of parts divided in hydraulic fracture (dimensionless). a channel with an opening equal to the initial fracture width,
Under a sparse distribution, the stress intensity factor for which is wf0 2 12. The two cases have the same trend. With an
a narrow fracture filled with proppants has two limits. On one increase in Rs1, the fracture conductivity first increases to a
side, in the proppant free part of the fracture, the stress certain value and then decreases. In other words, there is an
intensity factor is zero as the fracture faces have no cohesion optimal distance which has maximum conductivity after
(Neto and Kotousov 2013). This case will happen when there proppant embedment and deformation. This is because
is no proppant or highly compressible proppant. On the other increasing the proppant distance ratio has two different
side, when the whole fracture is continuously propped by effects on conductivity: (1) a reduction in proppant deforma-
incompressible particles, the stress intensity factors at the tip tion and embedment due to more proppants supporting the
of the fracture have a certain value, K0. The hydraulic fracture closure pressure, which decreases the loss of conductivity,
is divided in to NHF parts. Then K0 of each part can be and (2) an increase in the loss of conductivity due to a

606
J. Geophys. Eng. 14 (2017) 599 J Guo et al

Figure 18. Effect of rock elastic modulus on conductivity. (a) the


Figure 17. Effect of closure pressure on conductivity. (a) the
conductivity increases with the increased rock elastic modulus; (b)
conductivity decreases with the increased closure pressure; (b) the
the optimal distance ratio shows a slight decrease with an increase in
optimal distance ratio shows a slight increase with an increase in
the rock elastic modulus.
closure pressure.

at different Rs1. As shown in figure 17(a), the conductivity


reduced void volume which is occupied by more proppants.
decreases with the closure pressure for all the cases. As
The two mechanisms control the loss of conductivity,
shown in figure 17(b), the curve trends under four closure
resulting in the complex behavior shown in figure 16. As
pressures are slightly different, especially at the inflection
seen from this simulation, when the proppant distance ratio is
point. With an increase in the closure pressure, the inflection
less than 0.2 and 0.4 for the two cases, respectively, the
point moves to the right, meaning that Rs1 with maximum
reduction in proppant deformation and embedment mainly
conductivity increases. In other words, with an increase in the
controls the change in conductivity. When the proppant
closure pressure, the proppant packs should be designed to be
distance ratio increases and exceeds a certain value, the
closer to each other in order to achieve the maximum
reduction of a void volume primarily controls the change in
conductivity. Under a high closure pressure, the conductivity
conductivity. The curve trend is the same as the calculations
of a smaller Rs1 reduces to almost zero. This type of proppant
done by Khanna et al (2012), where the conductivity is
placement will be closed due to proppant embedment and
computed by an Ansys CFX with a monolayer of proppants.
deformation.
A reasonable agreement exists between the model developed
in this paper and their model.
The two cases have the same trend and, therefore, in the 3.2.3. Effect of rock elastic modulus. Figure 18 shows a
following sections, the basic parameters from Lacy et al relationship between conductivity and rock elastic modulus
(1998), shown in table 2, are used to study the effect of (E2) at different Rs1. As shown in figure 18(a), with an
different influencing factors. If not specified otherwise, the increase in E2, the conductivity increases, and the incremental
closure pressure is 30 MPa and Rs2=1. amplitude decreases. When E2 is large, a further increase in
E2 has little effect on conductivity. As shown in figure 18(b),
with an increase in E2, the inflection point moves to the left,
3.2.2. Effect of closure pressure. Figure 17 shows a which means that Rs1 with maximum conductivity decreases.
relationship between conductivity and closure pressure (Pc) Essentially, with an increase in E2, the proppant packs can be

607
J. Geophys. Eng. 14 (2017) 599 J Guo et al

Figure 19. Effect of proppant elastic modulus on conductivity. (a) Figure 20. Effect of proppant concentration on conductivity. (a) the
the conductivity increases with the increased proppant elastic conductivity increases with the increased proppant concentration; (b)
modulus; (b) the optimal distance ratio shows a slight decrease with the optimal distance ratio does not show an obvious change with the
an increase in proppant elastic modulus. variation of proppant concentration.

compression decreases, so the fracture width can be better


designed more diffusively to each other; with an increase in maintained.
E2, the ability to resist embedding increases, while the
deformation of rock under compression also decreases, so the 3.2.5. Effect of proppant concentration. Parameter J reflects
fracture width can be better maintained. Thus, the maximum the proppant concentration in a fracture. Figure 20 shows a
fracture conductivity can be achieved with a larger distance relationship between conductivity and proppant concentration
between proppant packs. (J) at different Rs1. As shown in figure 20(a), with an increase
in J, the conductivity, as well as the incremental amplitude,
increases. The results demonstrate that the relationship is a
3.2.4. Effect of proppant elastic modulus. Figure 19 shows a power law. As shown in figure 20(b), with an increase in J,
relationship between conductivity and proppant elastic the inflection point does not move obviously, which means
modulus (E1) at different Rs1. As shown in figure 19(a), that the Rs1 with maximum conductivity remains the same for
with an increase in E1, the conductivity increases and the the four cases. The proppant concentration does not show a
incremental amplitude decreases. The trend is the same as for discernable influence on the optimal distance.
E2 in figure 18(a). When E1 is large, a further increase in E1
has little effect on conductivity. As shown in figure 19(b), 3.2.6. Effect of poroelastic effect. During hydraulic
with an increase in E1, the inflection point moves to the left, fracturing, the fluid pressure near the wellbore is high,
which means that Rs1 with maximum conductivity decreases. while it is small near the tip. The two faces of the hydraulic
In other words, with an increase in E1, the proppant packs can fracture are not parallel to each other. The fracture width
be designed more diffusively to each other, because with an reduces from wellbore to the fracture tip along the
increase in E1, the deformation of proppants under longitudinal direction. The initial fracture width was

608
J. Geophys. Eng. 14 (2017) 599 J Guo et al

Figure 21. Comparison of parallel faces and elliptical shape. Fracture


conductivity is not a constant in an elliptical shape.

calculated by an analytical model (PKN) (Perkins and


Kern 1961, Nordgren 1972). It is assumed that the dynamic
fracture width calculated using the PKN model is completely
filled with proppants. Values for embedment and conductivity
were obtained by the model proposed in this paper.
Figure 21 shows a relationship between conductivity and
position along the fracture length direction at different
fracture shapes. The conductivity in parallel fractures is a
constant, while it decreases along the fracture length direction
in the elliptical shape. Generally, fracture conductivity is a
constant for a single fracture in a reservoir simulation.
Figure 21, however, shows that the possible actual case is
totally different from a constant. This difference may have an Figure 22. Effect of pressure drop on conductivity. (a) the
impact on simulated well productivity. conductivity decreases with the increased pressure drop; (b) the
In the production process, after the fluid diffusion from optimal distance ratio does not show an obvious change with the
variation of pressure drop.
the porous matrix to the fracture and then to the wellbore, the
poroelastic effect occurs. On the one hand, the pore pressure
decrease and the effective closure pressure increase will
increase the embedment depth. On the other hand, fluid flow with consideration of the sparse distribution of proppant
within the fracture causes a pressure drop (Δptt) along the packs.
fracture length direction. In order to better illustrate the effect (2) With an increase in the proppant distance ratio, the
of poroelastic, an elliptical shape was used. fracture conductivity first increases to a certain value
Because there is pressure drop inside the fracture. The and then decreases. An optimum proppant distance ratio
fluid pressure at the tip is higher than that at the near wellbore maintains the maximum conductivity after proppant
region. Therefore, the effective stress increases from the embedment and deformation.
fracture tip to near the wellbore region. With an increased (3) A change in the optimum proppant distance ratio
pressure drop, the effective stress near the wellbore increases. increases with closure pressure and decreases with the
Therefore, the conductivity near the wellbore is more rock elastic modulus and proppant elastic modulus.
sensitive to the pressure drop, as shown in figure 22(a). There is no obvious change with the variation of
However, the varied amplitude is limited, which does not proppant concentration.
show an effect on the optimal distance (figure 22(b)). (4) The stress intensity factor at the fracture tip increases
with the increased proppant distance ratio.

4. Conclusions Acknowledgments
Based on the study, the following conclusions can be drawn. The authors are grateful for the financial support of National
Natural Science Foundation of China (No. 51374178),
(1) New mathematical models have been derived to National Natural Science Foundation of China (No.
calculate changes in fracture width and conductivity 51604232), the Young Scholars Development Fund of SWPU

609
J. Geophys. Eng. 14 (2017) 599 J Guo et al

(201599010084) and NSERC/AIEES/Foundation CMG and Conference and Exhibition (San Antonio, TX, 5–8 October)
AITF Chairs (iCORE), and National Science and Technology (Tulsa, OK: Society of Petroleum Engineers) SPE-38590-MS
Major Project (No. 2016ZX05002-002). Lacy L L, Rickards A R and Bilden D M 1998 Fracture width and
embedment testing in soft reservoir sandstone SPE Drill
Completion 13 25–9
References Li K W, Gao Y P, Lyu Y C and Wang M 2014 New Mathematical
models for calculating proppant embedment and fracture
conductivity SPE J. 20 496–507
Awoleke O O, Zhu D and Hill A D 2016 New propped-fracture- Lu C, Guo J C, Wang W Y, Deng Y and Liu D F 2008 Experimental
conductivity models for tight gas sands SPE J. 21 1508–17 research on proppant embedment and its damage to fractures
Clark P E and Guler N 1983 Prop transport in vertical fractures: conductivity Nat. Gas Ind. 28 99–101
settling velocity correlations SPE/DOE Low Permeability Gas Mobbs A T and Hammond P S 2001 Computer simulations of
Reservoirs Symposium (Denver, CO, 14–16 March) (Tulsa, proppant transport in a hydraulic fracture SPE Prod. Facil. 16
OK: Society of Petroleum Engineers) SPE-11636-MS 112–21
D’Huteau E et al 2011 Open channel fracturing—a fast track to Mollanouri Shamsi M M, Farhadi Nia S and Jessen K 2015
production Oilfield Rev. Autumn 2011 23 Conductivity of proppant-packs under variable stress
Dontsov E V and Peirce A P 2014 Slurry flow, gravitational settling conditions: an integrated 3D discrete element and lattice
and a proppant transport model for hydraulic fractures J. Fluid boltzman method approach SPE Western Regional Meeting
Mech. 760 567–90 (Garden Grove, CA, 27–30 April) (Tulsa, OK: Society of
El-M.Shokir E M and Al-Quraishi A A 2009 Experimental and Petroleum Engineers) SPE-174046-MS
numerical investigation of proppant placement in hydraulic Neto L B, Khanna A and Kotousov A 2015 Conductivity and
fractures Pet. Sci. Technol. 27 1690–703 performance of hydraulic fractures partially filled with
Fredd C N, McConnell S B, Boney C L and England K W 2001 compressible proppant packs Int. J. Rock Mech. Min. 74 1–9
Experimental study of fracture conductivity for water- Neto L B and Kotousov A 2013 Residual opening of hydraulic
fracturing and conventional fracturing applications SPE J. 6 fractures filled with compressible proppant Int. J. Rock Mech.
288–98 Min. 61 223–30
Guo J C and Liu Y X 2012 Modeling of proppant embedment: Nordgren R P 1972 Propagation of a vertical hydraulic fracture SPE
elastic deformation and creep deformation SPE International J. 12 306–14
Production and Operations Conference & Exhibition (Doha, Perkins T K and Kern L R 1961 Widths of hydraulic fractures J. Pet.
Qatar, 14–16 May) (Tulsa, OK: Society of Petroleum Technol. 13 937–49
Engineers) SPE-157449-MS Sanematsu P et al 2015 Image-based Stokes flow modeling in bulk
Guo J C, Lu C, Zhao J Z and Wang W Y 2008 Experimental proppant packs and propped fractures under high loading
research on proppant embedment J. China Coal Soc. 33 661–4 stresses J. Petrol. Sci. Eng. 135 391–402
(in Chinese) Volk L J, Raible C J, Carroll H B and Spears J S 1981 Embedment
Hartley R and Bosma M 1985 Fracturing in chalk completions of high strength proppant into low-permeability reservoir rock
(includes associated papers 14483 and 14682) SPE J. Pet. SPE/DOE Low Permeability Gas Reservoirs Symposium
Technol. 37 73–9 (Denver, CO, 27–29 May) (Tulsa, OK: Society of Petroleum
Huitt J L and McGlothlin B B Jr 1986 The propping of fractures in Engineers) SPE-9867-MS
formations susceptible to propping-sand embedment Drilling Wells A A 1965 Notched bar tests, fracture mechanics and the brittle
and Production Practice (New York, 1 January) (Washington, strengths of welded structures Br. Weld. J. 12 2–13
DC: American Petroleum Institute) API-58-115 Wen Q Z, Zhang S C, Wang L, Liu Y S and Li X P 2007 The effect
Jin Z H, Johnson S E and Fan Z Q 2010 Subcritical propagation and of proppant embedment upon the long-term conductivity of
coalescence of oil-filled cracks: getting the oil out of low- fractures J. Petrol. Sci. Eng. 55 221–7
permeability source rocks Geophys. Res. Lett. 37 L01305 Yu W and Sepehrnoori K 2014 Simulation of gas desorption and
Johnson K L 1987 Contact Mechanics (Cambridge: Cambridge geomechanics effects for unconventional gas reservoirs Fuel
University Press) pp 92–3 116 455–64
Khanna A, Kotousov A, Sobey J and Weller P 2012 Conductivity of Zhang J J, Ouyang L C, Zhu D and Hill A D 2015 Experimental and
narrow fractures filled with a proppant monolayer J. Petrol. numerical studies of reduced fracture conductivity due to proppant
Sci. Eng. 100 9–13 embedment in the shale reservoir J. Petrol. Sci. Eng. 130 37–45
Lacy L L, Rickards A R and Ali S A 1997 Embedment and fracture Zou Y, Ma X, Zhang S, Zhou T, Ehlig-Economides C and Li H 2015
conductivity in soft formations associated with HEC, borate The origins of low-fracture conductivity in soft shale
and water-based fracture designs PE Annual Technical formations: an experimental study Energy Technol. 3 1233–42

610

You might also like