You are on page 1of 18

Journal of Petroleum Science and Engineering 158 (2017) 380–397

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Analytical model for production performance analysis of multi-fractured


horizontal well in tight oil reservoirs
Jinghao Ji, Yuedong Yao *, Shan Huang, Xiongqiang Ma, Shuang Zhang, Fengzhu Zhang
College of Petroleum Engineering, China University of Petroleum, Beijing 102249, China

A R T I C L E I N F O A B S T R A C T

Keywords: The combination of horizontal well drilling and reservoir stimulating via multiple fracturing has been proved to
Tight oil reservoirs be an effective technology to exploit low permeability unconventional tight oil reservoirs. Through multiple
Multi-fractured horizontal well fracturing, the very-complex fracture network called the stimulated reservoir volume (SRV) around primary
Production performance analysis
fractures can be generated, which is expected to affect well performance significantly. The pores in tight oil
Non-Darcy flow
reservoirs with multi-fractured horizontal wells (MFHWs) can be roughly divided into three major types: matrix,
Stress sensitivity
natural fractures and hydraulic fractures, which lead to multi-scaled flow during the development of tight oil
reservoirs. Furthermore, non-Darcy flow and stress sensitivity make the fluid flow in such reservoirs become more
complicated. So it is very difficult to accurately determine the productivity of MFHWs in these reservoirs. Various
analytical models have been established to analyze the transient performance of MFHW fast and accurately.
However, the typical seepage characteristics in tight oil reservoirs are seldom considered in these models, such as
non-Darcy flow and stress sensitivity.
This paper presents a new analytical model for MFHW in tight oil reservoirs, where the reservoir is subdivided
into five continuous flow regions: the unstimulated reservoir volume (USRV) regions 1 and 2, the stimulated
reservoir volume (SRV) regions 3 and 4, and the hydraulic fracture (HF) region 5. The effects of non-Darcy flow
and stress sensitivity are both considered. Laplace transformation, perturbation method and Stehfest numerical
inversion are employed to solve the model comprehensively. The model solution is verified with analytical and
numerical methods. The effects of related influential parameters on well production performance are investigated.
Field example is conducted to show the applicability of the new model, history matching curve and results are
obtained according to real production data. The research results summarized in this paper can provide some
significance for production performance analysis of MFHW in tight oil reservoirs.

1. Introduction analyzing the actual performance of MFHW in many unconventional


reservoirs (Zhao, 2012; Wang et al., 2014). The volume containing
Since horizontal wells with massive hydraulic fractures made the fracture networks which improve well performance is defined as SRV
exploitation of Bakken tight oil reservoirs in North America got a great (Mayerhofer et al., 2010), the reservoir that is not affected by fracturing
success (Liang et al., 2013), many countries and scholars began to pay is unstimulated reservoir volume (USRV).
more attention to the related theories of tight reservoirs. Due to the low Production performance analysis is an important tool in recognizing
porosity and low permeability of tight oil reservoirs, the fluid flowing formation properties and determining productivity of MFHW in uncon-
resistance is very large, resulting in low productivity of wells with con- ventional reservoirs. Due to the complexity of SRV, the existing con-
ventional developing method. MFHW has been proved to be an effective ventional models which ignored the SRV (Horne and Temeng, 1995;
method to exploit this kind of reservoirs. Multiple fracturing can not only Raghavan et al., 1997; Zerzar and Bettam, 2004; Al Rbeawi and Djebbar,
create several high conductivity primary fractures, but also activate 2012) are not able to predict well performance accurately. To deal with
natural fractures causing them open and connect with each other, and this challenge to characterize SRV, the methods in previous literatures
eventually generate large complex fracture networks (Maxwell et al., can be categorized into two major types: numerical simulation methods
2002; Clarkson, 2013; Zhou et al., 2014), which is a critical concept for and analytical/semi-analytical methods.

* Corresponding author.
E-mail address: 2015212172@student.cup.edu.cn (Y. Yao).

http://dx.doi.org/10.1016/j.petrol.2017.08.037
Received 17 September 2016; Received in revised form 22 June 2017; Accepted 14 August 2017
Available online 18 August 2017
0920-4105/© 2017 Elsevier B.V. All rights reserved.
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Nomenclature η Pressure conductivity coefficient, μm2 ⋅MPa=mPa⋅s


l Half-width of FSRV, m
pm , pf Pressure of matrix and fracture, MPa s Half-width of PSRV, m
t Time, h v Seepage velocity, m/h
qF Single fracture production rate, m3/d LR Reservoir length, m
nF Fracture number LH Horizontal well length, m
B Oil volume factor L The distance from the first fracture to the last fracture, m
μ Oil viscosity, mPa⋅s Lref Reference length, m
xF Fracture half-length (FSRV half-length), m WR Reservoir width, m
wF Fracture width, m h Formation thickness, m
Km , Kf Permeability of matrix and fracture, μm2
Subscript
Kref Reference permeability, μm2
D Dimensionless
ϕm , ϕf Porosity of matrix and fracture
m Matrix system
R matrix block radius, m
f Fracture system
λ Interporosity flow coefficient in SRV
F Hydraulic fracture
ω Storativity ratio in SRV
ref Reference
Ctm , Ctf Total compressibility coefficient of matrix and fracture,
1/MPa Superscript

CL Compressibility coefficient of oil, 1/MPa Laplace domain
λm Threshold pressure gradient, MPa/m
γ Permeability modulus, 1/MPa

(1) Numerical simulation methods: lots of work has been done to response and production performance of MFHW. Brown et al. (2009)
investigate the performance of MFHW by using numerical reservoir utilized the tri-linear concept of Lee and Brockenbrough (1986) and
simulators (Mayerhofer et al., 2006; Cipolla et al., 2009; Cipolla et al., proposed a “tri-linear flow” model (the model divided the reservoir into
2010; Weng et al., 2011; Wang et al., 2014), the fractures in SRV were three regions: USRV region, SRV region and HF region) to study the
simulated by high permeable refined grid cells and the effects of fracture transient pressure response of MFHW. This model was verified by
network complexity and conductivity on well performance were dis- comparing the results with the semi-analytical solution of Medeiros et al.
cussed in these studies. However, grid refinement can result in much (2008). According to the result of micro-seismic monitoring, the size of
computational time which limits the application of the numerical SRV around primary fractures is limited (Maxwell and Cipolla, 2011; Lin
methods. What's more, these methods can't reveal the microscopic et al., 2014), Stalgorova and Mattar (2012a, 2012b) proposed a “five-
percolation mechanisms. (2) Analytical/semi-analytical methods: to linear” model (extended model of “tri-linear flow” model) to simulate a
overcome the shortage of numerical simulation methods, analytical/ fracture that is surrounded by a stimulated region of limited extent which
semi-analytical methods are more attractive due to their accuracy and can be used for more-complex fracture network within SRV. Hereafter,
computational efficiency. Considering that rectangular drainage area the “linear flow” models were applied to shale gas reservoirs and the
better conforms to field practice and linear flow is the primary flow effects of ad-desorption and diffusion were considered (Ozkan et al.,
regime during the whole fractured horizontal well production stage, 2011; Brohi et al., 2011; Apaydin et al., 2012; Zhang et al., 2015a, 2016).
various “linear flow” models were proposed to analyze the pressure In addition, Yuan et al. (2015) established a new “multi-linear flow”

Fig. 1. Schematic of a MFHW in a closed box-shaped tight oil reservoir.

381
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Fig. 2. Schematic of the five-linear flow model for MFHW.

2015; Guo et al., 2016).


Table 1 All the previous work mentioned above is meaningful to better un-
Definitions of dimensionless variables.
derstand the transient performance of MFHW with SRV. It can be noted
p p
Dimensionless pressure pjD ¼ pi p
i j
wf
; j ¼ 1; 2; 3; 4; 5 that with the assumption of linear flow, the seepage model of MFHW are
Dimensionless production rate of single HF qFD ðtD Þ ¼ 1:84210
3
qF ðtÞμB
Kref hðpi pwf Þ
easy to establish and solve. What's more, the “linear flow” method can
Dimensionless time tD ¼
3:6ηref K
t (where ηref ¼ ðϕCtrefÞ μ) account for finite-conductivity HF without using the discretizing and
L2ref ref
numerical methods. Among these literatures, however, the unique flow
Dimensionless distance xD ¼ Lxref , yD ¼ Lyref , r3mD ¼ rR3m3 , r4mD ¼ rR4m4
mechanisms of non-Darcy flow and stress-sensitivity in tight oil reser-
Dimensionless width of HF wD ¼ LwrefF
voirs were not taken into consideration, which might significantly affect
Dimensionless half-length/width of FSRV xFD ¼ LxrefF , lD ¼ Lrefl
η
well production performance analysis. So they are not suitable to
Dimensionless diffusivity ηjD ¼ η j ; j ¼ 1; 2; 3; 4; 5
ref describe the fluid flow for MFHW in tight oil reservoirs. In view of the
Interporosity coefficient L2ref L2ref
λ3 ¼ KK3m
3fi R2
, λ4 ¼ KK4m
4fi R24
deficiency in the predecessor's research, a new analytical model for
3

Storativity ratio ω3 ¼ ðϕCt Þ


ðϕCt Þ3f
,
ðϕCt Þ4f
ω4 ¼ ðϕCt Þ
production performance analysis of MFHW in tight oil reservoirs is
3f þðϕCt Þ3m 4f þðϕCt Þ4m
proposed in this paper. Our model is based on five-linear flow regions
Dimensionless conductivity of HF CFD ¼ K5i wF
Kref Lref
including the USRV region, the SRV region and the HF region, the effect
Dimensionless Pseudo-TPG λD ¼ CL λm Lref
Dimensionless permeability modulus γ D ¼ ðpi  pwf Þγ
of non-Darcy flow caused by threshold pressure gradient (TPG) in the
Mobility ratio K =μ USRV region and the stress-sensitivity in the SRV region and HF region
M34 ¼ K4fi
3fi =μ
are both taken into consideration. Laplace transformation, perturbation
method are employed to solve the model, the analytical solution are then
inverted to real time domain with the algorithm proposed by Stehfest
(1970). We validate our model and a field example is conducted to show
the applicability of the new model. Finally, the related influential pa-
model on subdividing the reservoir into seven flow regions to account for rameters which influence the well production performance are investi-
different fracture intensity within fracture networks, which is more in gated to better understand the production performance of MFHW in tight
accordance with the real field geological situations. Another widely oil reservoirs.
studied models are “radial composite flow” models in which the basic
assumption is that SRV has a circle high permeable region around the 2. Physical model and assumptions
wellbore. Zhao et al. (2014) used a circle region to characterize SRV and
proposed a composite model to investigate the transient pressure and rate Fig. 1 shows a schematic of a MFHW in a closed box-shaped tight oil
behavior of MFHW in unconventional reservoirs. Similarly, rate transient reservoir. Due to symmetry, only half of one fracture stage is enough to
analysis for MFHW in tight oil reservoirs considering SRV was conducted derive flow equations as sketched in Fig. 2, the y-axes represents the
by Jiang et al. (2014). By incorporating the unique flow mechanisms of horizontal well direction and the x-axes represents the HF direction,
shale gas reservoirs, some improved models for MFHW in shale gas res- similar to the existing “linear flow” models, our equations are modified
ervoirs were proposed (Zhang et al., 2015b; Zeng et al., 2015; Su et al., from “five-linear” flow model, taking TPG and stress sensitivity into

Fig. 3. Schematic of fracture and matrix system.

382
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Fig. 4. Schematic of shape factors of different HFs.

Fig. 5. Comparison of our results with that in literature (Stalgorova and Mattar, 2012a).

consideration. Regions 1–5 represent flow in the HF, the SRV, and the 8
USRV regions, respectively. The USRV region (regions 1–2) in which < 3:6K ðgradðpÞ  λm Þ gradðpÞ > λm
>
most of the natural fractures are closed and have poor connectivity can be v¼ μ (1)
>
:
considered as single-porosity media, taking into account the TPG to 0 gradðpÞ  λm
describe non-Darcy flow. The SRV region (regions 3–4) can be treated as
dual-porosity system and finite-conductivity HF region (region 5) is In above equation, λm is the Pseudo-TPG.
treated as single-porosity media, while the permeability stress sensitivity In the SRV region and HF region, fracture permeability decreases as
cannot be neglected in regions 3–5. formation pressure depletes, which subsequently affects well perfor-
In the USRV region, reservoir properties are not affected by frac- mance. To account for the stress sensitivity in the SRV and HF regions, a
turing, the formation has tiny pore throat with ultra-low permeability, stress-dependent permeability is adopted, following Kikani and Pedrosa
one of the typical flow characteristics is the non-Darcy flow caused by (1991), the permeability modulus γ which is used to describe stress
TPG which might greatly affect well performance (Guo et al., 2012). sensitivity is defined as
There are different approaches to handle the non-Darcy flow in uncon-
1 ∂K
ventional reservoirs and lots of models have been proposed, however, γ¼ (2)
most of them are not practical to use due to their cumbersome calcula- K ∂p
tion. Therefore, the Pseudo-TPG approach is selected to describe the non- Micro-seismic data which represents the volumetric map of the extent
Darcy flow in this paper and expressed by the following equation (Prada of induced fracture network (Cipolla et al., 2008) shows that the SRV
and Civan, 1999; Song et al., 2007; Lei et al., 2008; Yin et al., 2013): always partially covers the area between primary HFs. Since region 3 is
farther away from HF than region 4, region 4 is defined as the fully

383
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Fig. 6. Simulator built by Eclipse to simulate production from tight oil reservoir.

(5) Oil flow in all five regions are approximate to be linear flow, both
Table 2
Basic parameters of the numerical model.
rock and fluid are slightly compressible and the compression co-
efficient is constant. The oil can only flow to wellbore through
Parameters Value Parameters Value
primary fractures.
Reservoir length, LR , m 580 Oil viscosity, μ, mPa⋅s 1.3
Reservoir width, WR , m 620 Horizontal well length, LH , m 400 It is worth noting that this paper chooses Pseudo-TPG approach to
Formation thickness, h, m 5 HF half Length, xF , m 110
USRV permeability, K1 ; K2 , mD 0.1 HF width, wF , m 0.005
describe non-Darcy flow in tight oil reservoirs. In fact, the Pseudo-TPG
Initial permeability of PSRV, 2 Initial permeability of HF, K5i , mD 20000 exaggerates the impact of non-Darcy flow. On the other hand, same
K3i , mD permeability modulus is used to characterize the stress sensitivity of SRV
Initial permeability of FSRV, 2 HF porosity, ϕ5 0.20 and HF for getting analytical solution of the proposed model, this may be
K4i , mD
inappropriate in some other cases. These are two limitations of the model
Porosity of USRV, ϕ1 ; ϕ2 0.15 HF number, nF 3
Porosity of SRV, ϕ3 ; ϕ4 0.20 Half- width of FSRV, l, m 100 and it is necessary to conduct some further researches considering the
Compressibility of oil, CL , 1/ 0.001 Half- width of PSRV, s, m 0 above issues.
MPa
Compressibility of rock, CR , 1/ 0.0005 Initial formation pressure, pi , MPa 25
MPa 3. Mathematical model
Threshold pressure gradient, 0 Bottom hole pressure, pwf , MPa 15
λm , MPa/m In order to facilitate the derivation of mathematical model, the
Oil volume factor, B 1.2 Permeability modulus of SRV and HF, 0.04 following dimensionless variables are defined in Table 1:
γ, 1/MPa
In the following, we present the mathematical model and explain the
derivation of solutions in different regions. The dual-porosity formula-
stimulated reservoir volume (FSRV) which has higher permeability, re- tion represents the SRV region, whereas the single-porosity formulation
gion 3 is treated as the partially stimulated reservoir volume (PSRV). The represents the USRV region and HF region. The model representing flow
location of the boundary shared by regions 3 and 4 can vary from near HF in different region is coupled by imposing the flux and pressure conti-
to y ¼ ye , which represents the effectiveness of induced fracture network. nuity across interfaces between the regions.
In order to make the problem more attractable, some other assumptions
are as follows:
3.1. Flow in the USRV region
(1) The MFHW producing at a constant bottom hole pressure (BHP) is
located centrally in a box-shaped tight oil reservoir with closed 3.1.1. Oil flow in region 1
outer boundaries, as shown in Fig. 1. The formation has the uni- In this region, we assume one-dimensional linear flow in the x-di-
form thickness h and the initial formation pressure is pi . rection, additionally, Pseudo-TPG is used to describe the non-Darcy flow,
(2) All the primary fractures are assumed to have identical properties which can be expressed as
and are equally distributed along the horizontal well, the fractures  
is symmetrical against well-axis and penetrate the formation 3:6K1 ∂p1
v1 ¼  λm (3)
entirely. μ ∂x
(3) No-flow boundary conditions are assumed between adjacent The diffusivity equation for flow in region 1 can be given by
fractures to simulate fracture interference, which means no fluids
flow at the middle of adjacent fractures (y ¼ ye ). ∂2 p 1 ∂p1 ϕ1 μCt1 ∂p1
(4) It is single-phase isothermal flow in the reservoir. Gravity, capil-  λm CL ¼ (4)
∂x2 ∂x 3:6K1 ∂t
lary force and the flow pressure loss along the horizontal well are
neglected. Incorporating the initial and boundary conditions, we can obtain the
dimensionless seepage model of region 1 with the dimensionless vari-
ables defined in Table 1, which can be described as

384
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Fig. 7. Comparison between analytical results and numerical solutions.

8 2
Table 3
> ∂ p2D ∂p2D 1 ∂p2D
>
> 2  λD ¼
Model parameters of the reservoir and well-1.
>
> ∂x ∂x η2D ∂tD
>
> D D
Parameters Value Parameters Value >
>
< p2D ðtD ¼ 0Þ ¼ 0
Reservoir length, LR , m 1320 Oil viscosity, μ, mPa⋅s 1.414  (6)
Reservoir width, WR , m 700 Horizontal well length, LH , m 1178 >
> ∂p 
> 2D 
> ¼0
Formation thickness, h, m 5 *HF half Length, xF , m 130 >
> ∂x
>
> D xD ¼xeD
Matrix permeability, 0.28 HF width, wF , m 0.003 >
: 
K1 ; K2 ; K3m ; K4m , mD p2D jxD ¼xFD ¼ p3fD xD ¼xFD
*Initial fracture permeability of 3 *Initial permeability of HF, K5i , mD 10000
PSRV, K3fi , mD
*Initial fracture permeability of 10 HF porosity, ϕ5 0.20
FSRV, K4fi , mD 3.2. Flow in the SRV region
Porosity of USRV, ϕ1 ; ϕ2 0.12 HF number, nF 10
Porosity of SRV, ϕ3ðf þmÞ ; ϕ4ðf þmÞ 0.15 *Half- width of FSRV, l, m 30
The dual-porosity formulation is adopted to describe flow in the SRV
Compressibility of matrix, Ctm , 1/ 0.0019 *Half- width of PSRV, s, m 35.4
MPa region, the oil flow in fracture system follows Darcy's law. To account for
Compressibility of fracture, Ctf , 1/ 0.0045 Initial formation pressure, pi , MPa 13.9 the stress sensitivity of fracture permeability, a stress-dependent
MPa permeability is adopted in the fracture system. According to the defini-
*Threshold pressure gradient, λm , 0.012 Bottom hole pressure, pwf , MPa 1.8 tion of permeability modulus in Eq. (2), we can get
MPa/m
Oil volume factor, B 1.19 *Permeability modulus of SRV and 0.04
K 1 p
HF, γ, 1/MPa ∫ Ki dK ¼∫ pi γdp (7)
K

Where pi is the initial formation pressure, Ki is the permeability at initial


condition, Eq. (7) can also be written as
8 ∂2 p ∂p1D 1 ∂p1D
>
>
1D
 λD ¼
>
> ∂x2D ∂x η K ¼ Ki eγðpi pÞ
>
> D 1D ∂tD (8)
>
>
>
< p1D ðtD ¼ 0Þ ¼ 0 Eq. (8) is the relationship that is used to describe stress-dependent
 (5)
>
> ∂p  permeability.
> 1D 
> ¼0
> ∂xD
>
>
> xD ¼xeD
3.2.1. Oil flow in region 3
>
: 
p1D j xD ¼xFD¼ p4fD  xD ¼xFD
With the assumption of linear flow in the y-direction and taking the
stress sensitivity into consideration, the equation for fracture system
3.1.2. Oil flow in region 2 including interporosity flux and supplement flux can be obtained
Assuming that one-dimensional linear flow in the x-direction and as follows
following similar derivation process like region 1, the dimensionless "  2 #
seepage model of region 2 is given by K3fi γðpi p3f Þ ∂2 p3f ∂p3f ϕ3f Ct3f ∂p3f
e þ γ þ q23 þ q3mf ¼ (9)
μ ∂y2 ∂y 3:6 ∂t

Where q23 represents the supplement flow term from region 2 to region 3,

385
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Fig. 8. Matching result of production rate for well-1.

q3mf denotes the interporosity flow term from matrix to fracture. Substituting Eqs. (11) and (12) into Eq. (9) and incorporating the
Due to the extreme low permeability, pseudo-steady state is difficult initial and boundary conditions, we can obtain the dimensionless
to reach, unsteady state interporosity flow is considered which is seepage model of region 3 as
reasonable for tight oil formations. De Swaan, 1976 model is applied to For the matrix system:
describe the oil flow in matrix, matrix blocks is assumed as spherical 8  
(Fig. 3) for convenience, and yield > 1 ∂ 2 ∂p3mD ð1  ω3 Þ ∂p3mD
>
> r ¼ ð0  r3mD  1Þ
>
> r 2
∂r 3mD
∂r λ3 η3D ∂tD
8   >
> 3mD 3mD 3mD
1 ∂ 2 ∂p3m ϕ μCt3m ∂p3m >
>
>
> ¼ 3m ð0  r3m  R3 Þ < p3mD ðr3mD ; 0Þ ¼ 0
>
> r3m ∂r3m
2 r3m
∂r 3:6k3m ∂t
>
> 3m  (13)
>
> >
> ∂p3mD 
>
< p3m ðr3m ; 0Þ ¼ pi >
> ¼0

>
>
> ∂r3mD r3mD ¼0
(10) >
>
>
>
> ∂p3m  :
>
> ∂r3m  ¼0 p3mD jr3mD ¼1 ¼ p3fD
>
>
>
>
r3m ¼0
: For the fracture system:
p3m jr3m ¼R3 ¼ p3f

8 "  2 #   
>
2
γ D p3fD ∂ p3fD ∂p3fD K2 ∂p2D  ∂p3mD  ω3 ∂p3fD
>
> e  γ þ þ G   3λ ¼
>
> ∂y2 D
∂yD K3fi xFD ∂xD  3
∂r3mD r3mD ¼1 η3D ∂tD
>
> x ¼x
>
>
D D FD

>
< p3fD ðtD ¼ 0Þ ¼ 0
 (14)
>
> γD p3fD ∂p3fD 
>
> ¼0
>
>
e
∂yD yD ¼yeD
>
>
>
>  
:
p3fD yD ¼lD ¼ p4fD yD ¼lD

Where q3mf can be expressed as λ L


In above equation, G ¼ pimprefwf .

3 K3m ∂p3m 
q3mf ¼ (11)
R3 μ ∂r3m r3m ¼R3 3.2.2. Oil flow in region 4
With the assumption of linear flow in the y-direction and following
The supplement flow term q23 from region 2 to region 3 is given by similar derivation process like region 3, the dimensionless seepage model
  to describe the oil flow in region 4 is given by
K2 1 ∂p2 
q23 ¼  λm  (12) For the matrix system:
μ xF ∂x x¼xF

386
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Table 4 which is given by


Reservoir and hydraulic fracture properties.

Parameters Value Parameters Value K4fi γ ðpi p4f Þ 2 ∂p4f 
q45 ¼ e (18)
Reservoir length, LR , m 1800 Oil viscosity, μ, mPa⋅s 5
μ wF ∂y y¼wF
2
Reservoir width, WR , m 600 Horizontal well length, LH , m 1400
Formation thickness, h, m 10 HF half Length, xF , m 100 Substituting Eq. (18) into Eq. (17) and incorporating the initial and
Matrix permeability, 0.1 HF width, wF , m 0.01 boundary conditions, we can obtain the dimensionless seepage model of
K1 ; K2 ; K3m ; K4m , mD region 5, which is described by
Initial fracture permeability of 2 Initial permeability of HF, K5i , 20000
PSRV, K3fi , mD mD 8 "  2 # 
HF porosity, ϕ5 >
>
2
γ D p5D ∂ p5D ∂p5D 2K4fi γD p4fD ∂p4fD  1 ∂p5D
Initial fracture permeability of 20 0.25 >
> e 2  γ þ e  wD ¼ η ∂t
FSRV, K4fi , mD >
> ∂x
D
∂x K w
5i D ∂y 5D
>
>
D D D yD ¼
2
D
Porosity of USRV, ϕ1 ; ϕ2 0.14 HF number, nF 8 >
>
>
Porosity of SRV, ϕ3ðf þmÞ ; ϕ4ðf þmÞ 0.20 Half- width of FSRV, l, m 20 < p5D ðtD ¼ 0Þ ¼ 0
>

Compressibility of matrix, Ctm , 0.00075 Half- width of PSRV, s, m 80
> ∂p5D 
1/MPa > eγD p5D
> ¼0
Compressibility of fracture, Ctf , 0.001 Initial formation pressure, pi , MPa 25 >
>
> ∂xD xD ¼xFD
>
> 
1/MPa >
> 
Threshold pressure gradient, λm , 0.02 Bottom hole pressure, pwf , MPa 15 > p5D ðxD ¼ 0Þ ¼ 1; eγ D p5D ∂p5D 
>
: ¼
π
qFD ðtD Þ
MPa/m ∂xD x ¼0 CFD
D
Oil volume factor, B 1.2 Permeability modulus of SRV and 0.02
HF, γ, MPa1 (19)
Matrix block radius of FSRV, R4 , 5 Matrix block radius of PSRV, R3 , 10
m m 4. Model solution
Reference length, Lref , m 30 Reference permeability, Kref , mD 20
Reference porosity, ϕref 0.2 Reference compressibility, Ctref , 0.00075
Note that the differential equations of region 1 and region 2 are
1/MPa
linear, which can be solved directly. However, because of the depen-
dence of permeability of SRV and HF on pore pressure, the equations of
regions 3–5 are strongly nonlinear. In order to eliminate the nonlinearity,
8   the Pedrosa's substitution (Pedrosa, 1986) is used with the expression as
> 1 ∂ 2 ∂p4mD ð1  ω4 Þ ∂p4mD
>
> r ¼ ð0  r4mD  1Þ
>
> r 2
∂r 4mD
∂r λ4 η4D ∂tD 1  
>
> 4mD 4mD 4mD
pjD ¼  ln 1  γ D ξjD j ¼ 3; 4; 5
>
> γD
(20)
< p4mD ðr4mD ; 0Þ ¼ 0
 (15)
>
>
> ∂p4mD  Performing a parameter perturbation in γ D by defining the
> ¼0
>
>
> ∂r4mD r4mD ¼0 following series
>
>
:
p4mD jr4mD ¼1 ¼ p4fD ξjD ¼ ξjD0 þ γ D ξjD1 þ γ 2D ξjD2 þ ⋯ j ¼ 3; 4; 5 (21)
For the fracture system:

8 "
>  2 #   
>
>
2
γ D p4fD ∂ p4fD ∂p4fD K1 ∂p1D  ∂p4mD  ω4 ∂p4fD
>
> e  γ þ þ G   3λ ¼
>
> 2
∂y ∂x  4
∂r  η
∂y 4D ∂tD
D
> K 4fi x 4mD
>
>
D D FD D xD ¼xFD r4mD ¼1
>
>
< p4fD ðtD ¼ 0Þ ¼ 0
  (16)
>
> 
γ D p4fD ∂p4fD  1 γD p3fD ∂p3fD 
>
> e ¼ e
>
> ∂yD yD ¼lD M34 ∂yD yD ¼lD
>
>
>
> 
>
>
: p4fD y ¼wD ¼ p5D jyD ¼wD
D 2 2

1
¼ 1 þ γ D ξjD þ γ 2D ξ2jD þ ⋯ j ¼ 3; 4; 5 (22)
λm Lref 1  γ D ξjD
In above equation, G ¼ pi pwf .

1   1
3.3. Flow in the HF region  ln 1  γ D ξjD ¼ ξjD þ γ D ξ2jD þ ⋯ j ¼ 3; 4; 5 (23)
γD 2
3.3.1. Oil flow in region 5 Considering that γ D is always small, the zero-order perturbation can
Flow in the HF region is also assumed to be linear flow in the x-di- meet the accuracy requirement (Wang, 2014; Chen et al., 2016). Based
rection. Taking into account the effect of stress sensitivity, the governing on the above analysis, the solutions of regions 3–5 can be obtained after
equation for flow in HF can be obtained as follows adopting perturbation technique to the seepage models.
"  2 #
K5i γðpi p5 Þ ∂2 p5 ∂p5 ϕ Ct5 ∂p5
e þγ þ q45 ¼ 5 (17) 4.1. Regions 1 and 2 solutions
μ ∂x2 ∂x 3:6 ∂t
Taking Laplace transformation with respect to tD to invert the seepage
Where q45 represents the supplement flow term from region 4 to region 5, models of regions 1–2 into Laplace domain, we can get

387
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

respectively in Laplace domain and we can get the solutions as follows


8 ∂2 p ∂p u
>
> 1D
 λD 1D  p ¼0  
>
> ∂x2D ∂xD η1D 1D
>
>  rr21 er1 ðxD xeD Þ þ er2 ðxD xeD Þ
<  
∂p1D  (24) p1D ¼  p4fD xD ¼xFD (26)
> ¼0
>
>
> ∂xD xD ¼xeD  rr21 er1 ðxFD xeD Þ þ er2 ðxFD xeD Þ
>
>
: 
p1D j ¼ p4fD 
xD ¼xFD xD ¼xFD
 
8 ∂2 p  mm21 em1 ðxD xeD Þ þ em2 ðxD xeD Þ 
> 2D ∂p u ¼  p3fD xD ¼xFD
>
>  λD 2D  p ¼0 p2D (27)
> ∂x2D
> ∂xD η2D 2D  mm21 em1 ðxFD xeD Þ þ em2 ðxFD xeD Þ
>
< 
∂p2D  (25) pffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffi

> ¼0
> ∂xD xD ¼xeD λD þ η4u þλ2D λD  η1D þλD
4u 2
>
> Where r1 ¼ 1D
, r2 ¼ ,
>
> pffiffiffiffiffiffiffiffiffiffi
ffi pffiffiffiffiffiffiffiffiffiffi
ffi2 2
: 
p2D jxD ¼xFD ¼ p3fD xD ¼xFD λD þ η þλD λD  η þλD
4u 2 4u 2

m1 ¼ 2
2D
, m2 ¼ 2
2D
.

Where u is the Laplace parameter, then the models can be solved

Fig. 9. Effect of TPG on production performance.

388
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

4.2. Regions 3 and 4 solutions 8


>  
>
> ∂2 ξ4D0 K1 ∂p1D G 
>
> þ þ  f4 ðuÞξ4D0 ¼ 0
After zero-order perturbation substitution and Laplace trans- >
>
> ∂y2D K4fi xFD ∂xD u xD ¼xFD
formation, the seepage models of regions 3–4 can be transformed as >
<  
∂ξ4D0  1 ∂ξ3D0  (29)
8 2   > ¼
>
>
∂ ξ3D0 K2 ∂p2D G  > ∂yD yD ¼lD M34 ∂yD yD ¼lD
>
>
> þ þ  f3 ðuÞξ3D0 ¼ 0 >
>
> ∂y2
>
> K3fi xFD ∂xD u xD ¼xFD >
>  
>
<
D
 : ξ4D0 yD ¼wD ¼ ξ5D0 yD ¼wD
>
∂ξ3D0 
2 2
(28) "
> ¼0 qffiffiffiffiffiffiffiffiffiffiffiffi
>
>
> ∂yD yD ¼yeD In Eqs. (28) and (29), f3 ðuÞ ¼ ω3 u
þ 3λ3 ð1ω3 Þu
coth
>
> ! # " η3D λ3 η3D
! #
>
:   qffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffi
ξ3D0 yD ¼lD ¼ ξ4D0 yD ¼lD ð1ω3 Þu
1 , f4 ðuÞ ¼ ωη4D
4u
þ 3λ4 ð1ω4 Þu
coth ð1ω4 Þu
 1 .
λ3 η3D λ4 η4D λ4 η4D


∂p2D 
 
Form Eq. (27), we can get ∂xD 
¼ β2 p3fD x ≈  β2 ξ3D0 xD ¼xFD ,
D ¼xFD
xD ¼xFD
m1 ðxFD xeD Þ m em2 ðxFD xeD Þ
whereβ2 ¼ m2 e  2
, defining K2
β
K3fi xFD 2
þ f3 ðuÞ ¼ α2 , the
m
m2 em1 ðxFD xeD Þ þem2 ðxFD xeD Þ
1

Fig. 10. Effect of stress sensitivity on production performance.

389
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

solution of region 3 can be solved as In above equation,


pffiffiffiffiffi pffiffiffiffiffi pffiffiffiffiffi i
cosh α2 ðyD  yeD Þ  K2 G K2 G αo sinh α1 ðyD  lD Þ þ cosh α1 ðyD  lD Þ
ξ3D0 ¼ pffiffiffiffiffi ξ4D0 yD ¼lD  þ A¼ hpffiffiffiffiffi i hpffiffiffiffiffi  lD (32)
cosh α2 ðlD  yeD Þ K3fi xFD α2 u K3fi xFD α2 u αo sinh α1 wD 2  lD þ cosh α1 wD 2
= =

 (30)
∂p1D 
   hpffiffiffiffiffi i
Similarly, ∂xD 
¼ β1 p4fD  xD ¼xFD
≈  β1 ξ4D0  xD ¼xFD
, where αo sinh α1 yD  wD 2 =
i
D ¼x
r2 er1 ðxFD xeD Þ r2 er2 ðxFDxx eD ÞFD
β1 ¼   , defining K1
β
K4fi xFD 1
þ f4 ðuÞ ¼ α1 and B¼ hpffiffiffiffiffi i hpffiffiffiffiffi  lD (33)
r2
r
1
er1 ðxFD xeD Þ þer2 ðxFD xeD Þ αo sinh α1 wD 2  lD þ cosh α1 wD 2
= =

pffiffiffiffi pffiffiffiffiffi
α
αo ¼ M34 p2ffiffiffi
α1
ffi tanh½ α2 ðyeD  lD Þ, the solution of region 4 can be solved as
4.3. Region 5 solution
 
 K1 G K1 K2 G
ξ4D0 ¼ A ξ5D0 yD ¼wD  þB  After zero-order perturbation substitution and Laplace trans-
2 K4fi xFD α1 u K4fi xFD α1 K3fi xFD α2 u
formation, the seepage model of region 5 can be transformed as
K1 G
þ
K4fi xFD α1 u
(31)

Fig. 11. Effect of fracture number on production performance.

390
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Fig. 12. Effect of fracture conductivity on production performance.

8 2  pffiffiffiffiffi hpffiffiffiffiffi i pffiffiffiffiffi hpffiffiffiffiffi i


>
> ∂ξ 2K4fi ∂ξ4D0  u β4 ¼ αo α1 cosh α1 wD 2  lD  α1 sinh α1 wD 2  lD
= =
> 5D0
> þ  wD  η ξ5D0 ¼ 0
>
> ∂x 2
K w ∂y hpffiffiffiffiffi i hpffiffiffiffiffi i
>
> D 5i D D yD ¼ 2 5D
αo sinh α1 wD 2  lD þ cosh α1 wD 2  lD
>
= =
>
< 

∂ξ5D0 
¼0 (34) (36)
>
>
> ∂xD xD ¼xFD
>
>  pffiffiffiffiffi
>
> i
>
> 1  eγD ∂ξ5D0  π αo α1
>
: ξ5D0 ðxD ¼ 0Þ ¼ ; ¼ q ðuÞ C¼ hpffiffiffiffiffi i hpffiffiffiffiffi  lD (37)
uγ D ∂xD xD ¼0 CFD FD αo sinh α1 wD 2  lD þ cosh α1 wD 2
= =


Taking the derivative of region 4 solution, 2K 1 β4 2K4fi C
defining α4 ¼ K5i w4fiD β4 þ ηu and β5 ¼ K5i2K G
þ K1
     5D wD xFD α1 u K5i wD K4fi xFD α1
∂ξ4D0   K1 G K1
¼ β4 ξ5D0 yD ¼wD  K3fiKxFD G
þC u , we can get the single HF production rate in Laplace domain,
2

∂yD yD ¼wD 2 K4fi xFD α1 u K4fi xFD α1
α2
2
 which is expressed as
K2 G
 (35)
K3fi xFD α2 u

Where,

391
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Fig. 13. Effect of FSRV half-length on production performance.

pffiffiffiffiffi pffiffiffiffiffi   
1  eγD CFD α4 tanh α4 xFD CFD β pffiffiffiffiffi
qFD ðuÞ ¼ ⋅  pffiffiffiffi5ffi tanhð α4 xFD Þ min i;N2
γD πu π α4 X K 2 ð2KÞ!
N
N þi
(38) Vi ¼ ð1Þ 2   (40)
iþ1
N
2
 K !K!ðK  1Þ!ði  KÞ!ð2K  iÞ!
K¼ 2
4.4. Well production solution

According to Eq. (38), we can get the single HF dimensionless pro- As can be seen in Fig. 2, the flow rate of a single HF element is
duction rate in real-time domain by Stehfest (1970) algorithm numeri- determined by its dominated drainage area that could described by a
cal inversion shape factor, which can be expressed as

  xe
ln 2 X
N
ln 2 δ¼ (41)
qFD ðtD Þ ¼ Vi qFD i (39) ye
tD i¼1 tD
As for the whole rectangular drainage area in Fig. 4, different HFs
along the horizontal well have different shape factors, which could be
classified as two main types:

392
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Fig. 14. Effect of FSRV half-width on production performance.

For interior HFs: 5. Discussion of the results

WR =2 WR 5.1. Model validation and field example


δin ¼ ¼ (42)
L=2ðnF  1Þ L=ðnF  1Þ
5.1.1. Analytical validation
For exterior HFs:
Previously, the “five-linear flow” model has been validated through
WR =2 WR numerical simulation method (Stalgorova and Mattar, 2012a). The
δex ¼ ¼ (43) dimensionless BHP has been obtained in that literature, according to Van
ðLR  LÞ=2 LR  L
Everdingen and Hurst (1949), the corresponding dimensionless produc-
Incorporating with fracture number nF , the well production rate and tion rate can be obtained as follows
cumulative production at a constant BHP of MFHW satisfies the following
equations (Meyer et al., 2010): 1
qD ¼ (46)
s2 pwD
qD ðtD Þ ¼ ðnF  1ÞqFD ðδin ; tD Þ þ qFD ðδex ; tD Þ (44)
To verify the accuracy of our model, a relatively simple case is con-
t ducted and comparison is made with the model proposed by Stalgorova
QD ðtD Þ ¼ ∫ 0D qD ðtD ÞdtD (45) and Mattar (2012a). In this paper, the fracture network around primary
fractures and the effects of TPG in USRV region and stress sensitivity in

393
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Fig. 15. Effect of SRV fracture permeability on production performance.

SRV/HF region are both considered. If some parameters are changed, the region. The other basic parameters are summarized in Table 2. We make
new proposed model can be simplified and converted to some other different permeability modulus γ ¼ 0 and γ ¼ 0:04 MPa1 in the com-
existing models. If we set λm ¼ 0, γ ¼ 0, f3 ðuÞ ¼ η3D
u
; f4 ðuÞ ¼ η4D
u
, this new parison. The numerical solutions is compared with our calculated results
model is similar with the model in literature (Stalgorova and Mattar, in Fig. 7, as can be seen, in spite of a little difference, the analytical so-
2012a). The basic input parameters and the comparison results are lutions obtained in this paper are consistent with the numerical solutions,
shown in Fig. 5, as can be seen, the single fracture production rate which indicates that our model is reliable.
calculated by our model is in good accordance with the existing model,
which verifies that our model is trustable. 5.1.3. Field example
To show the applicability of the new model proposed in this paper, oil
5.1.2. Numerical validation production rate data collected under constant BHP of well-1 from Car-
To further validate the analytical solution in this paper, a numerical dium tight oil formation (Clarkson and Pedersen, 2011) is provided here
model built by Eclipse numerical simulator to simulate production from to perform production performance analysis. It is a MFHW with 10
tight oil reservoir is applied. The global grid is 29  31  1 and each cell transverse hydraulic fractures and the average fracture spacing is 429ft
is 20  20  10 m (Fig. 6). Local grid refinement (LGR) method is utilized (130.76 m). Basic parameters can be obtained from the reservoir and
to model oil flow into HF. For convenience, the single porosity model fracturing information to conduct history matching. The basic input pa-
(f3 ðuÞ ¼ ηu ; f4 ðuÞ ¼ ηu ) is used and the pseudo-TPG value of 0 MPa/m is rameters are shown in Table 3 and the matching curve of field data is
3D 4D

assumed because TPG is not considered in Eclipse, the keywords shown in Fig. 8. The matching parameters are listed in Table 3, which are
“ROCKTAB” is applied to describe stress sensitivity effect in HF/SRV highlighted with the asterisk (*). It can be found from the matching curve

394
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Fig. 16. Effect of interporosity flow model on production performance.

(the red curve) that our model can achieve good agreement with the sensitivity into the production performance analysis of MFHW in tight oil
actual production data, which indicates that our model is a useful tool to reservoirs.
analyze the production performance of MFHW in tight oil reservoirs.
In order to show the impacts of non-Darcy flow and stress sensitivity 5.2. Sensitivity analysis
on production performance analysis, the matching curve are compared
with 3 different cases, namely, other parameters unchanged without both To better understand the effect of influential parameters on well
non-Darcy flow and stress sensitivity (case 1), other parameters un-
production performance, we conduct the sensitivity analysis of the pro-
changed without stress sensitivity (case 2), other parameters unchanged posed model. Table 4 presents the synthetic data, which gives the
without non-Darcy flow (case 3). Of course, production performance
reservoir properties and fracture properties, respectively. The key factors
analysis has more than one solutions and by changing other parameters, that influence well production performance including TPG, stress sensi-
the same match may be obtained in these three cases. But from the
tivity, fracture number, fracture conductivity, FSRV half-length, FSRV
compared curves, it clearly shows that non-Darcy flow and stress sensi- half-width, SRV fracture permeability and interporosity flow model are
tivity have big effects on regression analysis, which indicates that
discussed in detail.
ignoring the effects of non-Darcy flow and stress sensitivity may cause
unreasonable results in production performance analysis of MFHW in
5.2.1. Effect of TPG
tight oil reservoirs. From above analysis, the model in this paper is an Fig. 9 demonstrates the TPG on well production performance. Spe-
improved model that can taking the effects of non-Darcy flow and stress
cifically, Eq. (1) is a discriminant, thus there is a constant negative

395
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

outflow. Because of this, the dimensionless production rate will become when FSRV half-length reaches a certain value, the cumulative produc-
negative after a certain production time, which means the well will not tion increase trend will slows down, which is mainly caused by the
produce any more. The existence of TPG (non-Darcy flow) can signifi- fractures disturbance under the condition of limited spaces between
cantly make the production drop quicker compared with the situation adjacent HFs.
without TPG, while the early production performance remain un-
changed. From the compared curves in Fig. 9, we can find that the bigger 5.2.7. Effect of SRV fracture permeability
TPG which means the flow resistance in USRV is higher, the faster the As can be seen from Fig. 15, which shows the effect of SRV fracture
production rate will decline and the smaller the final cumulative pro- permeability on production performance of MFHW in tight oil reservoirs,
duction will be. when SRV fracture permeability increases, the flow resistance can be
evidently reduced. And hence, well production rate increases effectively.
5.2.2. Effect of stress sensitivity This phenomenon demonstrates that improving the fracture permeability
It is shown in Fig. 10 that stress sensitivity affects all the production in SRV region is very critical for the development of tight oil reservoirs.
periods. This is because hydraulic fractures and fracture networks (SRV)
are the primary channels for fluid to flow to the horizontal well, the 5.2.8. Effect of interporosity flow model
existence of stress sensitivity in the SRV and HF will lead to the lower Fig. 16 is the comparison of production performance curves of MFHW
well production rate compared with the situation without the effect of in tight oil reservoirs calculated by pseudo-steady interporosity flow
stress sensitivity. The increase of permeability modulus can significantly model (PSM) and unsteady interporosity flow model (USM). When
lower the well production rate and final cumulative production. calculating by using PSM (Warren and Root, 1963), f3 ðuÞ and f4 ðuÞ in In
Eqs. (28) and (29) are replaced by f3 ðuÞ ¼ λ3 ηλ3D
3 ð1ω3 Þu ω3 u
þð1ω3 Þu þ η3D and
5.2.3. Effect of fracture number
f4 ðuÞ ¼ λ4 ηλ4 ð1ω 4 Þu ω4 u
þð1ω4 Þu þ η . The production rate of USM is higher than that
The well production performance with different fracture number is 4D 4D

shown in Fig. 11. It shows that fracture number has significant effect on of PSM in the early production period, The reason is that matrix-fracture
early-mid production time, the number of fracture reflects the scale of oil exchange is more sensitive to pressure change in USM, which leads to
hydraulic fracturing, so the lager fracture number means a larger and oil supplement from matrix to fracture is more faster in USM. In the later
better seepage area for fluid to flow from reservoir to wellbore, which production periods, the production calculated by these two models be-
finally leads to higher cumulative well production. However, the increase comes identical. This also demonstrates that the USM is more appropriate
of fracture number will lead to a lower span between primary fractures in tight formations.
which indicates the stronger adjacent fractures interference, resulting in
the trend of cumulative production increase slows down. The result in- 6. Conclusions
dicates that there is an optimal fracture number under a
certain condition. This paper presents an analytical model for production analysis of
MFHW in tight oil reservoirs. The model is verified and a field example is
5.2.4. Effect of fracture conductivity conducted. By investigating the transient production behavior and
Fig. 12 indicates the effect of fracture conductivity on well production analyzing the effects of related influential parameters, the main conclu-
performance. It can be seen that fracture conductivity mainly influences sions of this paper are given in follows:
the well production in the early and intermediate production periods.
The lager value of fracture conductivity indicates the oil flow in HF is (1) The proposed model considering multi-scaled flow and complex
easier, so the bigger the fracture conductivity is, the higher position the flow mechanisms (non-Darcy flow and stress sensitivity) to
well production rate will be, which also benefits to final cumulative describe the flow of a MFHW in tight oil reservoirs is much better
production. However, when fracture conductivity becomes lager, the conforming to real situation.
cumulative production increase trend slows down. This is mainly because (2) TPG (non-Darcy flow) and stress sensitivity have significant in-
the limited capacity of fluid supplementary from matrix to fracture in fluences on transient production performance of MFHW in tight
tight oil reservoirs. Thus, facture conductivity should match the prop- oil reservoirs. Lower TPG and permeability modulus of SRV/HF
erties of a specific tight oil formation. can result in higher production rate.
(3) Fracture number, fracture half-length (FSRV half-length), fracture
5.2.5. Effect of FSRV half-length conductivity and FSRV half-width all have effects on well pro-
Fig. 13 shows the effect of the half-length of FSRV on well production duction performance, the increase of them can result in higher
performance. It is known that the increasing of FSRV half-length can cumulative production. However, the cumulative production in-
improve the seepage condition and increase the oil drainage area which crease trend would slow down when they reach to a certain level.
decreases the flow resistance for fluid to flow into wellbore. It can be seen (4) The results of this study can be helpful to better understand the
that the magnitude of FSRV half-length can significantly change the well flow mechanisms and transient production performance of MFHW
production behavior. With the increase of FSRV half-length, the well in tight oil reservoirs.
production rate will be lager which benefits to the final cumulative
production. However, it can also be observed that the trend of cumulative Acknowledgements
production increase slows down when FSRV half-length becomes lager.
Thus, there is an optimal FSRV half-length to get economical This research was supported by the National Basic Research Program
development. of China (2015CB250900), the Program for New Century Excellent Tal-
ents in University (Grant No.NCET-13-1030) and the National Natural
5.2.6. Effect of FSRV half-width Science Foundation of China (Grant No.40974055).
Fig. 14 compares the difference of production behavior of MFHW
with different FSRV half-width. When the other parameters keep con- References
stant, the lager half-width of FSRV indicates the effective stimulated
Al Rbeawi, S.J.H., Djebbar, T., 2012. Transient pressure analysis of a horizontal well with
reservoir volume is lager. Due to the higher permeability of FSRV, the multiple inclined hydraulic fractures using type-curve matching. In: SPE International
flow resistance near the wellbore area can be reduced. The lager value Symposium and Exhibition on Formation Damage Control. Society of Petroleum
the half-width is, the higher the well production rate and final cumula- Engineers.
tive production will be obtained at early production time. However,

396
J. Ji et al. Journal of Petroleum Science and Engineering 158 (2017) 380–397

Apaydin, O.G., Ozkan, E., Raghavan, R., 2012. Effect of discontinuous microfractures on Medeiros, F., Ozkan, E., Kazemi, H., 2008. Productivity and drainage area of fractured
ultratight matrix permeability of a dual-porosity medium. SPE Reserv. Eval. Eng. 15 horizontal wells in tight gas reservoirs. SPE Reserv. Eval. Eng. 11 (05), 902–911.
(04), 473–485. Meyer, B.R., Bazan, L.W., Jacot, R.H., et al., 2010. Optimization of multiple transverse
Brohi, I.G., Pooladi-Darvish, M., Aguilera, R., 2011. Modeling fractured horizontal wells hydraulic fractures in horizontal wellbores. In: SPE Unconventional Gas Conference.
as dual porosity composite reservoirs-application to tight gas, shale gas and tight oil Society of Petroleum Engineers.
cases. In: SPE Western North American Region Meeting. Society of Petroleum Ozkan, E., Brown, M.L., Raghavan, R., et al., 2011. Comparison of fractured-horizontal-
Engineers. well performance in tight sand and shale reservoirs. SPE Reserv. Eval. Eng. 14 (02),
Brown, M.L., Ozkan, E., Raghavan, R.S., et al., 2009. Practical solutions for pressure 248–259.
transient responses of fractured horizontal wells in unconventional reservoirs. In: SPE Pedrosa Jr., O.A., 1986, January. Pressure transient response in stress-sensitive
Annual Technical Conference and Exhibition. Society of Petroleum Engineers. formations. In: SPE California Regional Meeting. Society of Petroleum Engineers.
Chen, Z., Liao, X., Zhao, X., et al., 2016. Development of a trilinear-flow model for carbon Prada, A., Civan, F., 1999. Modification of Darcy's law for the threshold pressure gradient.
sequestration in depleted shale. SPE J. 21 (04), 1–13. J. Petrol. Sci. Eng. 22 (4), 237–240.
Cipolla, C.L., Warpinski, N.R., Mayerhofer, M.J., et al., 2008. The relationship between Raghavan, R.S., Chen, C.C., Agarwal, B., 1997. An analysis of horizontal wells intercepted
fracture complexity, reservoir properties, and fracture treatment design. In: SPE by multiple fractures. SPE J. 2 (03), 235–245.
Annual Technical Conference and Exhibition. Society of Petroleum Engineers. Song, F.Q., Jiang, R.J., Bian, S.L., 2007. Measurement of threshold pressure gradient of
Cipolla, C.L., Lolon, E., Mayerhofer, M.J., 2009. Reservoir modeling and production microchannels by static method. Chin. Phys. Lett. 24, 1995–1998.
evaluation in shale-gas reservoirs. In: International Petroleum Technology Stalgorova, E., Mattar, L., 2012a. Analytical model for history matching and forecasting
Conference. production in multifrac composite systems. In: SPE Canadian Unconventional
Cipolla, C.L., Lolon, E.P., Erdle, J.C., et al., 2010. Modeling well performance in shale-gas Resources Conference. Society of Petroleum Engineers.
reservoirs. SPE Reserv. Eval. Eng. 13 (4), 638–653. Stalgorova, E., Mattar, L., 2012b. Practical analytical model to simulate production of
Clarkson, C.R., 2013. Production data analysis of unconventional gas wells: review of horizontal wells with branch fractures. In: SPE Canadian Unconventional Resources
theory and best practices. Int. J. Coal Geol. 109, 101–146. Conference. Society of Petroleum Engineers.
Clarkson, C.R., Pedersen, P.K., 2011. Production analysis of Western Canadian Stehfest, H., 1970. Algorithm 368: numerical inversion of Laplace transforms. Commun.
unconventional light oil plays. In: Canadian Unconventional Resources Conference. ACM 13 (1), 47–49.
Society of Petroleum Engineers. Su, Y., Zhang, Q., Wang, W., et al., 2015. Performance analysis of a composite dual-
de Swaan, O.A., 1976. Analytic solutions for determining naturally fractured reservoir porosity model in multi-scale fractured shale reservoir. J. Nat. Gas Sci. Eng. 26,
properties by well testing. Soc. Petrol. Eng. J. 16 (03), 117–122. 1107–1118.
Van Everdingen, A.F., Hurst, W., 1949. The application of the Laplace transformation to Wang, H.T., 2014. Performance of multiple fractured horizontal wells in shale gas
flow problems in reservoirs. J. Petrol. Technol. 1 (12), 305–324. reservoirs with consideration of multiple mechanisms. J. Hydrol. 510, 299–312.
Guo, J., Zhang, S., Zhang, L., et al., 2012. Well testing analysis for horizontal well with Wang, H., Liao, X., Lu, N., et al., 2014. A study on development effect of horizontal well
consideration of threshold pressure gradient in tight gas reservoirs. J. Hydrodyn., Ser. with SRV in unconventional tight oil reservoir. J. Energy Inst. 87 (2), 114–120.
B 24 (4), 561–568. Warren, J.E., Root, P.J., 1963. The behavior of naturally fractured reservoirs. Soc. Petrol.
Guo, J., Wang, H., Zhang, L., 2016. Transient pressure and production dynamics of multi- Eng. J. 3 (03), 245–255.
stage fractured horizontal wells in shale gas reservoirs with stimulated reservoir Weng, X., Kresse, O., Cohen, C.E., et al., 2011. Modeling of hydraulic-fracture-network
volume. J. Nat. Gas Sci. Eng. 35, 425–443. propagation in a naturally fractured formation. SPE Prod. Oper. 26 (04), 368–380.
Horne, R., Temeng, K.O., 1995. Relative productivities and pressure transient modeling of Yin, D., Zhou, Y., Yuan, H., Zhang, C., 2013. The numerical simulation of non-darcy flow
horizontal wells with multiple fractures. In: Middle East Oil Show. Society of for yushulin low permeability oilfield. Int. J. Control Autom. 6 (5), 323–344.
Petroleum Engineers. Yuan, B., Su, Y., Moghanloo, R.G., et al., 2015. A new analytical multi-linear solution for
Jiang, R., Xu, J., Sun, Z., et al., 2014. Rate transient analysis for multistage fractured gas flow toward fractured horizontal wells with different fracture intensity. J. Nat.
horizontal well in tight oil reservoirs considering stimulated reservoir volume. Math. Gas Sci. Eng. 23, 227–238.
Probl. Eng. 2014. Zeng, H., Fan, D., Yao, J., et al., 2015. Pressure and rate transient analysis of composite
Kikani, J., Pedrosa Jr., O.A., 1991. Perturbation analysis of stress-sensitive reservoirs shale gas reservoirs considering multiple mechanisms. J. Nat. Gas Sci. Eng. 27,
(includes associated papers 25281 and 25292). SPE Form. Eval. 6 (03), 379–386. 914–925.
Lee, S.T., Brockenbrough, J.R., 1986. A new approximate analytic solution for finite- Zerzar, A., Bettam, Y., 2004. Interpretation of multiple hydraulically fractured horizontal
conductivity vertical fractures. SPE Form. Eval. 1 (01), 75–88. wells in closed systems. In: Canadian International Petroleum Conference. Petroleum
Lei, Q., Xiong, W., Yuan, J., Shusheng Gao, S., Wu, Y.S., 2008. Behavior of flow through Society of Canada.
low-permeability reservoirs. In: Europec/EAGE Conference and Exhibition. Society of Zhang, D., Zhang, L., Guo, J., et al., 2015a. Research on the production performance of
Petroleum Engineers. multistage fractured horizontal well in shale gas reservoir. J. Nat. Gas Sci. Eng. 26,
Liang, T., Chang, Y., Guo, X., et al., 2013. Influence factors of single well's productivity in 279–289.
the bakken tight oil reservoir, williston basin. Petrol. Explor. Dev. 40 (3), 383–388. Zhang, D., Zhang, L., Zhao, Y., et al., 2015b. A composite model to analyze the decline
Lin, M., Chen, S., Chen, Z., et al., 2014. Fractured reservoir modeling: effects of hydraulic performance of a multiple fractured horizontal well in shale reservoirs. J. Nat. Gas
fracture geometries in tight oil reservoirs. In: SPE/EAGE European Unconventional Sci. Eng. 26, 999–1010.
Resources Conference and Exhibition. Zhang, L., Gao, J., Hu, S., et al., 2016. Five-region flow model for MFHWs in dual porous
Maxwell, S.C., Cipolla, C.L., 2011. What does microseismicity tell us about hydraulic shale gas reservoirs. J. Nat. Gas Sci. Eng. 33, 1316–1323.
fracturing?. In: SPE Annual Technical Conference and Exhibition. Society of Zhao, G., 2012. A simplified engineering model integrated stimulated reservoir volume
Petroleum Engineers. (SRV) and tight formation characterization with multistage fractured horizontal
Maxwell, S.C., Urbancic, T.I., Steinsberger, N., et al., 2002. Microseismic imaging of wells. In: SPE Canadian Unconventional Resources Conference. Society of Petroleum
hydraulic fracture complexity in the Barnett shale. In: SPE Annual Technical Engineers.
Conference and Exhibition. Society of Petroleum Engineers. Zhao, Y.L., Zhang, L.H., Luo, J.X., et al., 2014. Performance of fractured horizontal well
Mayerhofer, M.J., Lolon, E.P., Youngblood, J.E., 2006. Integration of microseismic- with stimulated reservoir volume in unconventional gas reservoir. J. Hydrol. 512,
fracture-mapping results with numerical fracture network production modeling in the 447–456.
Barnett Shale. In: SPE Annual Technical Conference and Exhibition. Society of Zhou, W., Banerjee, R., Poe, B.D., et al., 2014. Semianalytical production simulation of
Petroleum Engineers. complex hydraulic-fracture networks. SPE J. 19 (01), 6–18.
Mayerhofer, M.J., Lolon, E., Warpinski, N.R., et al., 2010. What is stimulated reservoir
volume? SPE Prod. Oper. 25 (01), 89–98.

397

You might also like