You are on page 1of 17

Transport in Porous Media

https://doi.org/10.1007/s11242-022-01835-y

S.I.: GEOLOGICAL SYSTEMS WITH FRACTURED POROUS MEDIA

Modeling of Shale Gas Transport in Multi‑Scale Complex


Fracture Networks Considering Fracture Hits

Bin Li1

Received: 24 March 2022 / Accepted: 16 July 2022


© The Author(s), under exclusive licence to Springer Nature B.V. 2022

Abstract
Shale gas reservoir is a complex multi-scale system containing micro-nanopores and
micro-fractures. Understanding shale gas transport mechanism in fractured porous media
is important to predict shale gas production performance accurately. This paper established
a shale gas production prediction model, considering gas rarefaction effects, adsorption,
diffusion, and stress sensitivity. The variation in production and drainage patterns with pro-
duction time by multi-stage and multi-cluster fracturing considering fracture hit was stud-
ied by using this model. In addition, the influences of connecting hydraulic fractures, natu-
ral fracture conductivity, and stress sensitivity on shale gas production are discussed. When
the spacing between the connecting fractures exceeds 4 stages (176 m), the production of
the child well and parent well tends to be stable as the spacing between connecting hydrau-
lic fractures increases. The child well production decreases, and the parent well production
increases by considering fracture hits. The cumulative production of both parent and child
wells increases with the increase in natural fracture conductivity. The results show that the
production of parent and child wells considering stress sensitivity is, respectively, 14.31%
and 18.73% lower than that without considering stress sensitivity. The key findings of this
study can be expected to provide theoretical supports for the shale gas transport mecha-
nisms in fractured porous media.

Keywords Shale gas · Gas transport mechanism · Multi-scale fractures networks · Stress
sensitivity · Fracture hits

Abbreviations
De Effective diameter of nanopores for gas transport, m
kbulk Bulk diffusion apparent permeability, ­m2
­ 2
kf Permeability of the fractures, m
kk Knudsen diffusion apparent permeability, ­m2
km Permeability of shale matrix, ­m2
km0 Permeability at atmospheric pressure
ks Surface diffusion apparent permeability, ­m2
kvs Slip flow apparent permeability, ­m2

* Bin Li
libin3@cnooc.com.cn
1
China United Coalbed Methane Corporation Ltd, Beijing, China

13
Vol.:(0123456789)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
B. Li

M Gas molar mass, kg/mol


n Number of moles of gas, mol
na Adsorption amount, mol/kg
n0 Maximum adsorption capacity, mol/kg
p Pressure, Pa
pL Langmuir pressure, Pa
p0 Gas pressure at a certain time during depressurization, Pa
R Universal gas constant, Pa·m3/(mol·K)
T Temperature, K
Tc Critical temperature of gas, K
u Gas velocity in fractures, m/s
um Velocity, m/s
V Volume, ­m3
wf Fracture width, m
Z Gas compressibility factor
μ Gas viscosity, Pa·s
ξb Coefficient factor of adsorption gas for bulk gas transport
ξs Coefficient factor of adsorption gas for surface diffusion
ρg Gas density, kg/m3
ρr Shale density, kg/m3
φf Fracture porosity
φm Shale matrix porosity
φm0 Porosity at atmospheric pressure

1 Introduction

With the rapid development of the world economy, the demand for fossil energy is increas-
ing day by day. It is estimated that the world’s primary energy demand for fuel will
reach 1.715 million tons of oil equivalent by 2035, of which natural gas will account for
25%(Yang et al. 2019). Moreover, the demand for natural gas will increase to nearly 500
billion cubic meters by 2040(Wang et al. 2022). Unconventional natural gas resources,
especially shale gas, are abundant and have great development potential, playing a crucial
role in energy supply (Ran et al. 2022; Sun et al. 2022).
However, the extremely low porosity and permeability of shale greatly hinder shale gas
development. Hydraulic fracturing has become a key technology for the industrial develop-
ment of shale gas resources (Sharifi et al. 2021), that is, the injection of high-pressure fluid to
form artificial fractures in the reservoir and improve the reservoir permeability. The conven-
tional hydraulic fracturing method forms only one hydraulic fracture per stage. For tight gas
reservoirs, the stimulation of a single hydraulic fracture on production is temporary, which
cannot achieve the goal of economic and efficiency. Therefore, in order to greatly improve
single well production, increase stable production time, and improve economic benefits, shale
gas wells are often fractured by multi-stage and multi-cluster fracturing technology. Multi-
stage and multi-cluster fracturing technology can create multiple hydraulic fractures in each
stage and use interfracture interference to create complex fracture networks, improve near-
wellbore reservoir permeability, increase stimulation volume, and increase production (Wang
and Fidelibus 2021). In recent years, infill drilling generally produces a higher fracture hit
risk between infill well and its neighboring wells. Fracture hits in unconventional oil and gas

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Modeling of Shale Gas Transport in Multi‑Scale Complex Fracture…

reservoirs have attached much research interest. Therefore, it is necessary to study the multi-
stage fracturing production in multi-scale complex fracture networks considering fracture hits.
Production prediction of shale gas after multistage fracturing is a difficult research topic
at present. Analytical method, semi-analytical method and numerical simulation method are
often used to study the productivity and fracture parameter sensitivity of fractured horizontal
wells in tight reservoirs (Yang et al. 2018; Jiang and Younis 2015; Lee et al. 2001; Lee and
Lee 2015; Monteagudo and Firoozabadi 2004; Shi et al. 2021; Geng et al. 2018; Ning et al.
2020). However, the characteristic of fracture network formed by fracturing in shale reservoir
is complex, the flow mode of fluid in the stimulation area has not been accurately described
(Olorode et al. 2020), and the physical model corresponding to the actual fracturing situa-
tion has not been established for multi-stage and multi-cluster fracturing. The multi-stage and
multi-cluster fracturing of shale gas horizontal wells takes the perforation section as the center
to form multiple clusters of fractures, which are interwoven together to form a complex frac-
ture network system, rather than a single fracture. The flow pattern of shale gas in the stimula-
tion area is very complex.
Shale gas reservoir is a complex multi-scale system containing micro-nanopores, micro-
fractures, and fractures. Micropores exist in the matrix and are generally on the order of
nanometers in size and act as gas storage containers through adsorption(Ran et al. 2022).
The diffusion and flow of gas in micropores are mainly driven by a concentration difference.
Microcracks mainly store compressed gas, and the gas flow in microcracks is mainly driven
by the pressure difference (Ran et al. 2022). Due to nanoscale phenomena and complex frac-
ture networks in shale, the gas flow mechanism in shale is different from that in conventional
reservoirs. Under the condition of the original formation, shale gas mainly occurs in the shale
matrix as an adsorption state. When the pressure of the shale gas reservoir falls below the
critical desorption pressure, shale gas is desorbed from the surface of the shale matrix, and the
original dynamic balance is broken. As methane molecules are continuously desorbed from
the surface of shale matrix, the molecular concentration near the desorption region is higher
than that in the non-desorption region, resulting in a difference in gas concentration, which
causes the diffusion of methane molecules from the high concentration region to the low con-
centration region. Methane molecules through desorption and diffusion flow through matrix
pores. Because the pore size of shale matrix is similar to the mean free path of gas molecules,
gas molecules are not only affected by viscous force but also collide with rock walls, resulting
in a slip effect. The gas flow mechanism in the reservoir can be divided into continuous flow,
slip flow, transition flow, and free molecular flow. Therefore, multiple migration mechanisms
must be considered in the model to predict shale gas production accurately. A single constitu-
tive equation cannot accurately describe the flow characteristics of shale gas, and it is a chal-
lenge to quantitatively describe the gas flow in shale reservoirs after multi-stage and multi-
cluster fracturing.
To accurately predict the productivity of multi-stage and multi-cluster fracturing consider-
ing fracture hits in shale gas horizontal well, a productivity prediction model for shale gas
horizontal well was established considering various migration mechanisms and complex frac-
ture morphology. Based on verifying the accuracy of the model, the influence of geological
and connecting hydraulic fractures on production is deeply analyzed.

2 Model Description

The governing equation for gas flow in the shale matrix is (Lee et al. 1966):

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
B. Li

( ) [( ) ]
) 𝜕 𝜌g 𝜑m 𝜕 1 − 𝜑m 𝜌r na M
(1)
(
−∇ ⋅ 𝜌g um = +
𝜕t 𝜕t
where ρg (kg/m3) is the gas density; φm is the shale matrix porosity; ρr (kg/m3) is the shale
density; na (mol/kg) is adsorption amount; and M (kg/mol) is gas molar mass. um (m/s) is
the velocity, which can be described by Darcy’s law:
km
um = − ∇p (2)
𝜇
where km (m2) is the permeability of the shale matrix; and μ (Pa·s) is the gas viscosity.
The real gas state equation can be formulated as:
pV = ZnRT (3)
3
where p (Pa) is the pressure; V ­(m ) is the volume; n (mol) is the number of moles of gas; R
(Pa·m3/(mol·K)) is the universal gas constant; T (K) is temperature. Z is the gas compress-
ibility factor, which can be written as (Mahmoud 2014):

Z = 0.044Tr2 − 0.164Tr + 0.702e−2.5Tr p2r − 5.524e−2.5Tr pr + 1.15 (4)

T
Tr =
Tc (5)

p
pr =
pc (6)

where pc (Pa) is the critical pressure of gas; Tc (K) is the critical temperature of gas.
Gas viscosity can be estimated as:
𝜌g c�
(7)

𝜇 = a� eb ( 1000 ) 10−7

(9.379 + 0.01607M)(1.8T)1.5
a� = (8)
209.2 + 19.26M + 1.8T

986.4
b� = 3.448 + + 0.01M (9)
1.8T

c� = 2.447 − 0.2224b� (10)


The permeability of shale matrix can be calculated as (Wang et al. 2019, 2020):
km = kvs + kbulk + kk + ks (11)

De 2
1 8Kn
kvs = 𝜉b (1 + 2𝛼Kn)(1 + ) (12)
(1 + Kn) 32 1 − 2bKn

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Modeling of Shale Gas Transport in Multi‑Scale Complex Fracture…


1 De 𝜇 8ZRT 1 p 𝜕Z
kbulk = 𝜉b ( − ) (13)
(1 + Kn) 3 p 𝜋M Z 2Z 2 𝜕p


Kn De 𝜇 8ZRT 1 p 𝜕Z
kk = 𝜉b ( − ) (14)
(1 + Kn) 3 p 𝜋M Z 2Z 2 𝜕p

𝜇RT 0 1 p 𝜕Z
ks = 𝜉s Ds Ca max [ − ] (15)
p ZpL + p Z(ZpL + p) 𝜕p

where ξb is the coefficient factor of adsorption gas for bulk gas transport; kvs ­(m2) is the slip
flow apparent permeability; kbulk ­(m2) is the bulk diffusion apparent permeability; kk ­(m2)
is the Knudsen diffusion apparent permeability; ξs is coefficient factor of adsorption gas for
surface diffusion; ks ­(m2) is the surface diffusion apparent permeability; De (m) is the effec-
tive diameter of nanopores for gas transport(Wang et al. 2019).
The adsorption amount na can be determined by the Langmuir model (Langmuir 1918):
P
na = n0
pL + p (16)

where n0 (mol/kg) is the maximum adsorption capacity and pL (Pa) is the Langmuir
pressure.
The permeability and porosity of the shale matrix during the production process can be
calculated by (Dong et al. 2010; Wu et al. 2016):
pc − p −qp
( )
𝜑m = 𝜑m0 (17)
p0

( )−sk
pc − p
km = km0 (18)
p0

where φm0 is the porosity at atmospheric pressure; km0 is the permeability at atmospheric
pressure; pc (Pa) is the overburden pressure. p0 (Pa) is gas pressure at a certain time during
depressurization.
The governing equation for gas flow in fractures is:
( )
𝜕 𝜌g 𝜑f
(19)
( )
wf + ∇T ⋅ wf 𝜌g uf = 0
𝜕t
where φf is the fracture porosity and wf (m) is the fracture width; uf (m/s) is the gas velocity
in fractures, which can be described by Darcy’s law:
kf
uf = − ∇T p (20)
𝜇
where kf ­(m2) is the permeability of the fractures.
The porosity of the fractures is written as follows (Wang et al. 2020):

𝜑f = 𝜑f 0 e−cf (pini −p) (21)

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
B. Li

The permeability of fractures can be written as (Neto et al. 2015):


)3 ( )2
𝜑f 1 − 𝜑f 0
(
kf = k f 0 (22)
𝜑f 0 1 − 𝜑f

The width of fractures can be written as follows (Wang et al. 2020):


1 − 𝜑f 0
wf = wf 0 (23)
1 − 𝜑f

Based on the above numerical model, a two-dimensional shale gas multi-stage and
multi-cluster fracturing model was established considering the micro-nano characteristics
of shale gas. The size of the gas reservoir is 1600 m × 350 m. The hydraulic fractures in
each cluster are of unequal length, and the natural fractures are generated uniformly and
randomly. The reservoir is surrounded by no-flow boundary conditions, the bottom-hole
flow pressure of horizontal wells is constant, and gas flows into the wellbore only through
hydraulic fractures. The geometric model of the base case is shown in Fig. 1.
To establish a multi-stage and multi-cluster fracturing model considering fracture hits in
multi-scale complex fracture networks, the open-source software ADFNE is used to gen-
erate discrete fractures, and the PyGeoMesh is used to establish the reservoir domain as
the front-end for grid generation of the open-source software Gmsh. Finally, the commer-
cial software Comsol is used for numerical simulation. The simulation workflow is shown
in Fig. 2. PyGeoMesh is an automatic mesh generator for generic and multiscale porous
media. It uses the powerful Gmsh as the backend. Compared with the mesh generator that
comes with Comsol, the mesh generated by PyGeoMesh and Gmsh has better adaptability
to the meshing after considering discrete fractures. PyGeoMesh realizes the mesh optimi-
zation of the model after adding discrete fractures by setting the overall mesh size for the
whole domain and mesh size for fracture, trace, and well.

3 Results and Discussion

3.1 Model Validation

The production data of a field example from the Weiyuan Block shale are adopted to vali-
date the present model. The production process of this gas well lasted 1140 days. The lat-
eral length of this well is 1500 m, and it has 25 stimulated fracture stages. The reservoir
model with length and width of 1600 m × 350 m is set up, and the relevant parameters are

Fig. 1  Geology model of shale gas multi-stage and multi-cluster fracturing. It has 25 stages of hydraulic
fractures, and each stage has 3 clusters

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Modeling of Shale Gas Transport in Multi‑Scale Complex Fracture…

Fig. 2  Simulation workflow

shown in Table 1 and Table 2. The model’s reliability can be testified because the actual
cumulative gas production for 1140 days of the gas well reaches 76.4 × ­106 ­m3 (Xu et al.
2022), while the calculated cumulative gas production for 1140 days is 75.2 × ­106 ­m3, with
a relative error of 1.615%. The daily and cumulative production curves of a single well are
shown in Fig. 3. It can be seen that the daily production of a single well is the largest in the
early stage of production, and with the increase in production time, the production gradu-
ally decreases and eventually tends to be stable. During shale gas development, the mini-
mum pressure occurs around hydraulic fractures. Pressure waves are first generated within
the cluster and then gradually spread outwards. In the early stage of exploitation, due to the
communication between hydraulic fractures and matrix, shale gas around hydraulic frac-
tures is exploited faster, resulting in a faster rate of pressure drop between hydraulic frac-
tures. With the production time increase, the shale gas flowing from reservoir to hydraulic

Table 1  Reservoir parameters Parameter Value Parameter Value

P0 45 MPa n0 0.12 mol/kg


T 392.4 K φf0 0.032
Pw 5 MPa Lf 100 m
φm0 0.038 Wf0 0.3 mm
km0 447 × ­10–6 mD Kf0 750 mD
PL 8.5 MPa H 30 m

Table 2  Fracture parameters


Stage number Stage spacing (m) Cluster number Cluster spacing (m)

25 40 3 2

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
B. Li

fracture decreases, and the velocity of pressure wave propagation and pressure drop slows
down.
In order to investigate the impact on pressure response and well productivity for con-
necting hydraulic fractures, we set up a basic reservoir model with parent and child wells
based on the model for validation. The relevant parameters are consistent with Table 1 and
Table 2. When the parent well has been in production for 1000 days, the child well starts
to produce as shown in Fig. 4. After accumulative production of 2000 days, the cumula-
tive production of the parent well increases by 1.74% compared to the model for valida-
tion without child well. The reason for this is that there is mutual interference between the
wells.

3.2 Effect of Connecting Hydraulic Fractures

3.2.1 Number of Connecting Hydraulic Fractures

This section studies the influence of different numbers of connecting hydraulic fractures on
shale gas production. The simulation parameters are shown in Table 3, and the results are
shown in Fig. 5 and Fig. 6. It can be seen that with the increase in the number of fracturing
hits, the production of the child well decreases, the production of the parent well increases,
and the total production of the two wells decreases. The reason is that when two horizontal
wells are directly connected through hydraulic fractures, the increase in production of the
parent well is less than the decrease in the production of the child well. Compared with
no connecting fractures, the cumulative production of the parent well with 13 connecting
fractures increased by 5.73%, while the production of the child well decreased by 18.80%.
For the parent well, when the number of connected fractures increases to 7, the daily pro-
duction after 1000 days first increases and then decreases. This shows that the impact of
Table 3  Parameters with number of connecting hydraulic fractures
Number of connecting hydrau- Connecting hydraulic fracture Connecting fracture conductivity (mD·m)
lic fractures spacing (m)

2/4/5/7/13 1056/352/264/176/88 0.225

(a) Daily gas production (b) Cumulative gas production.

Fig. 3  Production of a base case

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Modeling of Shale Gas Transport in Multi‑Scale Complex Fracture…

(a) Without child well

(b) With child well

Fig. 4  Pressure contours for basic reservoir model at 2000 day

fracture hits on the production of the parent well and the child well is more obvious when
the child well is put into production at an early stage.

3.2.2 Connecting Hydraulic Fracture Spacing

This section studies the influence of different connecting hydraulic fracture spacing on
shale gas production. The simulation parameters are shown in Table 4, and the results are
shown in Fig. 7 and Fig. 8. Take two connecting hydraulic fractures as an example. The
distance between connecting hydraulic fractures is 0 m, which means there are no con-
necting fractures. The cumulative gas production of the parent and child wells increases
with the spacing increase. But the increase in parent and child wells production gradu-
ally slows down with the increasing interval spacing. When the spacing between the con-
necting fractures exceeds 4 stages, the production of parent and child wells tends to be
stable. The child well production is significantly more affected by connecting hydraulic
fracture spacing than the parent well. The reason is that when the interval is small, the
interaction between fractures is serious and the formation pressure drops greatly, resulting
in a decrease in the production of each fracture, as shown in Fig. 7. With the increase in
fracture spacing, the mutual interference between fractures weakens, and the production of
each fracture increases, but the increased amplitude becomes slow.

Table 4  Parameters with different connecting hydraulic fracture spacing


Number of connecting hydrau- Connecting hydraulic fracture Connecting fracture conductivity (mD·m)
lic fractures spacing (m)

2 0/44/88/176/352/704 0.225

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
B. Li

(a) 2 connecting hydraulic fractures

(b) 4 connecting hydraulic fractures

(c) 5 connecting hydraulic fractures

(d) 7 connecting hydraulic fractures

(e) 13 connecting hydraulic fractures

Fig. 5  Pressure contours with different numbers of connecting fractures at 2000 day

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Modeling of Shale Gas Transport in Multi‑Scale Complex Fracture…

(a) Daily gas production of parent well (b) Cumulative gas production of parent well

(c) Daily gas production of childwell (d) Cumulative gas production of child well

(e) Daily gas production of the two wells (f) Cumulative gas p roduction of the two wells

Fig. 6  Shale gas production with different numbers of connecting fractures

3.2.3 Connecting Fracture Conductivity

This section studies the influence of different connecting fracture conductivity on shale gas
production. The simulation parameters are shown in Table 5, and the results are shown in
Fig. 9. It can be seen that increasing the connecting fracture conductivity can improve the
cumulative production of child well and reduce the cumulative production of parent well
production. The reason is that the more clusters there are, the more reservoir volume can
be stimulation and more production can be produced in a single stage.

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
B. Li

(a) 0 m (b) 44 m

(c) 88 m (d) 176 m

(e) 352 m (f) 704 m

Fig. 7  Pressure contours with different connecting hydraulic fracture spacing at 2000 day

(a) Parent well (b) Child well

Fig. 8  Cumulative gas production with different numbers of connecting fractures at 2000 day

3.3 Effect of Geological Factors

3.3.1 Natural Fracture Conductivity

This section studies the influence of natural fracture conductivity on shale gas production
considering fracture hits. The simulation parameters are shown in Table 6, and the results
are shown in Fig. 10. It can be seen that natural fracture permeability greatly impacts
the productivity of horizontal well with fracture hits. Unlike the effect of the connecting
fracture conductivity, the cumulative production of both parent and child wells increases
with the increase in natural fracture conductivity. When the natural fracture conductivity
is low, the conductivity obviously affects productivity, and the cumulative gas produc-
tion increases significantly. When the natural fracture conductivity is high, the influence
of conductivity on productivity is gradually weakened, and the increase in cumulative
gas production is small. When the natural fracture conductivity increased from 0.005 to
5 mD, the cumulative production growth rate of the parent well decreased from 1.85 to

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Modeling of Shale Gas Transport in Multi‑Scale Complex Fracture…

Table 5  Parameters with different connecting fracture conductivity


Number of connecting hydrau- Connecting hydraulic fracture Connecting fracture conductivity (mD·m)
lic fractures spacing (m)

4 352 0.225/2.25/22.5/225

(a) Parent well (b) Child well

Fig. 9  Cumulative gas production with different connecting fracture conductivity

Table 6  Parameters with different fracture permeability and aperture


Natural fracture Number of con- Connecting hydraulic Connecting fracture conductivity (mD·m)
conductivity necting hydraulic fracture spacing (m)
(mD·m) fractures

0.005/0.05/0.5/5 4 352 0.225

0.95%. Moreover, the formation pressure is little affected by conductivity during produc-
tion, which is shown in Fig. 11. It also can be seen that with the increase in natural fracture
aperture, the cumulative gas production per well increases. The main reason is that the for-
mation conductivity is determined by the natural fracture permeability and aperture, mak-
ing it easier for shale gas to migrate in the reservoir.

3.3.2 Stress Sensitivity

This section studies the influence of stress sensitivity on shale gas production with fracture
hits. The simulation parameters are the same as the case of the previous section except for
the natural fracture conductivity. The cumulative production of both parent and child wells
after 2000 days with and without considering stress sensitivity is compared in Fig. 12.
It can be seen that shale gas production considering stress sensitivity is lower than that
without stress sensitivity for both parent and child wells, and the difference is significant.
The cumulative gas production of parent and child wells considering stress sensitivity at
2000 days is reduced by an average of 14.31% and 18.73% compared with the case without
considering stress sensitivity. And the impact of stress sensitivity on cumulative production

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
B. Li

Fig. 10  Cumulative production with different natural fracture conductivity

(a) 0.005 mD·m (b) 0.05 mD·m

(c) 0.5 mD·m (d) 5 mD·m

Fig. 11  Pressure contours at 2000 day

(a) Parent well (b) Child well

Fig. 12  Effect of the stress sensitivity on shale gas production (ss for stress sensitive and fh for fracture hits)

is more obvious in the case of fracture hits. The reason is that the apparent permeability is
greatly overestimated without considering stress sensitivity, resulting in higher daily pro-
duction. Therefore, the influence of stress sensitivity must be considered for multi-stage
and multi-cluster fracturing.

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Modeling of Shale Gas Transport in Multi‑Scale Complex Fracture…

4 Conclusions

This study aims to predict better the productivity of shale gas horizontal well consider-
ing fracture hits. The variation in production and formation pressure with production time
after multi-stage and multi-cluster fracturing is studied. In addition, the influences of the
number of connecting hydraulic fractures, connecting hydraulic fracture spacing, connect-
ing fracture conductivity, natural fracture conductivity, and stress sensitivity on shale gas
production are discussed. The following conclusions can be drawn in this work.
(1) With the increase in the number of fracturing hits, the production of the child well
decreases, the production of the parent well increases, and the total production of the two
wells decreases;
(2) When the spacing between the connecting fractures exceeds 4 stages, the production
of the child well and parent well tends to be stable;
(3) The cumulative production of the parent well decreases with the increase in the con-
necting fracture conductivity and increases with the increase in natural fracture conductiv-
ity. The cumulative production of child well increases with the conductivity of connecting
and natural fractures;
(4) The cumulative gas production of parent and child wells considering stress sensitiv-
ity at 2000 days is reduced by an average of 14.31% and 18.73% compared with the case
without considering stress sensitivity. And the impact of stress sensitivity on cumulative
production is more obvious in the case with fracture hits.

Supplementary Information The online version contains supplementary material available at https://​doi.​
org/​10.​1007/​s11242-​022-​01835-y.

Acknowledgements The authors would like to thank the Research on Key Technologies for CUCBM’s Pro-
duction of 6 Billion Cubic Meters (CNOOC-KJ 135 ZDXM 40) for financial support.

Funding This work has been funded by the Research on Key Technologies for CUCBM’s Production of 6
Billion Cubic Meters (CNOOC-KJ 135 ZDXM 40).

Conflict of interest The authors have not disclosed any competing interests.

References
Dong, J., Hsu, J., Wu, W., Shimamoto, T., Hung, J., Yeh, E., Wu, Y., Sone, H.: Stress-dependence of the
permeability and porosity of sandstone and shale from TCDP hole-A. Int. J. Rock Mech. Min. Sci. 47,
1141–1157 (2010)
Geng, L., Li, G., Wang, M., Li, Y., Tian, S., Pang, W., Lyu, Z.: A fractal production prediction model for
shale gas reservoirs. J. Nat. Gas Sci. Eng. 55, 354–367 (2018)
Jiang, J., Younis, R.M.: Numerical study of complex fracture geometries for unconventional gas reservoirs
using a discrete fracture-matrix model. J.Nat. Gas Sci. Eng. 26, 1174–1186 (2015)
Langmuir, I.: The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 40,
1361–1403 (1918)
Lee, S.J., Lee, K.S.: Performance of shale gas reservoirs with nonuniform multiple hydraulic fractures.
Energy Sour. Part A: Recovery, Util. Environ. Effects 37(13), 1455–1463 (2015)
Lee, A.L., Gonzalez, M.H., Eakin, B.E.: The viscosity of natural gases. J. Petrol. Technol. 18(997–991), 000
(1966)
Lee, S.H., Lough, M., Jensen, C.: Hierarchical modeling of flow in naturally fractured formations with mul-
tiple length scales. Water Resour. Res. 37, 443–455 (2001)

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
B. Li

Mahmoud, M.: Development of a new correlation of gas compressibility factor (Z-factor) for high pressure
gas reservoirs. J. Energy Res. Technol. 136, 012903 (2014)
Monteagudo, J.E.P., Firoozabadi, A.: Control-volume method for numerical simulation of two-phase immis-
cible flow in two- and three-dimensional discrete-fractured media: simulation of flow in fractured
media. Water Resour. Res. (2004). https://​doi.​org/​10.​1029/​2003W​R0029​96
Neto, L.B., Khanna, A., Kotousov, A.: Conductivity and performance of hydraulic fractures partially filled
with compressible proppant packs. Int. J. Rock Mech. Min. Sci. 74, 1–9 (2015)
Ning, X., Feng, Y., Wang, B.: Numerical simulation of channel fracturing technology in developing shale
gas reservoirs. J. Nat. Gas Sci. Eng. 83, 103515 (2020)
Olorode, O., Wang, B., Rashid, H.U.: Three-dimensional projection-based embedded discrete-fracture
model for compositional simulation of fractured reservoirs. SPE J. 25(04), 2143–2161 (2020)
Ran, L., Gaomin, L., Jinzhou, Z., Lan, R., Jianfa, W.: Productivity model of shale gas fractured horizontal
well considering complex fracture morphology. J. Petrol. Sci. Eng. 208, 109511 (2022)
Sharifi, M., Kelkar, M., and Karkevandi-Talkhooncheh, A. J. A. i. G.-E. R.: A workflow for flow simulation
in shale oil reservoirs: a case study in Woodford shale. Adv. Geo-Energy Res. 5, 365–375 (2021)
Shi, Y., Song, X., Song, G.: Productivity prediction of a multilateral-well geothermal system based on a
long short-term memory and multi-layer perceptron combinational neural network. Appl. Energy 282,
116046 (2021)
Sun, Y., Ju, Y., Zhou, W., Qiao, P., Tao, L., and Xiao, L. J. A. i. G.-E. R.: Nanoscale pore and crack evolu-
tion in shear thin layers of shales and the shale gas reservoir effect. Adv. Geo-Energy Res. 6, 221–229
(2022)
Wang, B., Fidelibus, C.: An open-source code for fluid flow simulations in unconventional fractured reser-
voirs. Geosciences 11(2), 106 (2021)
Wang, T., Tian, S., Li, G., Zhang, P.: Analytical model for real gas transport in shale reservoirs with surface
diffusion of adsorbed gas. Ind. Eng. Chem. Res. 58, 23481–23489 (2019)
Wang, T., Tian, S., Zhang, W., Ren, W., Li, G.: Production model of a fractured horizontal well in shale gas
reservoirs. Energy Fuels 35, 493–500 (2020)
Wang, L., Yao, Y., Wang, K., Adenutsi, C.D., Zhao, G., Lai, F.: Data-driven multi-objective optimization
design method for shale gas fracturing parameters. J. Nat. Gas Sci. Eng. 99, 104420 (2022). https://​doi.​
org/​10.​1016/j.​jngse.​2022.​104420
Wu, K., Chen, Z., Li, X., Guo, C., Wei, M.: A model for multiple transport mechanisms through nanopores
of shale gas reservoirs with real gas effect–adsorption-mechanic coupling. Int. J. Heat Mass Transf. 93,
408–426 (2016)
Xu, Y., Liu, X., Hu, Z., Shao, N., Duan, X., Chang, J.: Production effect evaluation of shale gas fractured
horizontal well under variable production and variable pressure. J. Nat. Gas Sci. Eng. 97, 104344
(2022)
Yang, Z.-Z., Yi, L.-P., Li, X.-G., He, W.: Pseudo-three-dimensional numerical model and investigation of
multi-cluster fracturing within a stage in a horizontal well. J. Petrol. Sci. Eng. 162, 190–213 (2018)
Yang, R., Hong, C., Huang, Z., Song, X., Zhang, S., Wen, H.: Coal breakage using abrasive liquid nitrogen
jet and its implications for coalbed methane recovery. Appl. Energy 253, 113485 (2019)

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the
author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is
solely governed by the terms of such publishing agreement and applicable law.

13
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center
GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers
and authorised users (“Users”), for small-scale personal, non-commercial use provided that all
copyright, trade and service marks and other proprietary notices are maintained. By accessing,
sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of
use (“Terms”). For these purposes, Springer Nature considers academic use (by researchers and
students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and
conditions, a relevant site licence or a personal subscription. These Terms will prevail over any
conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription (to
the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of
the Creative Commons license used will apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may
also use these personal data internally within ResearchGate and Springer Nature and as agreed share
it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not otherwise
disclose your personal data outside the ResearchGate or the Springer Nature group of companies
unless we have your permission as detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial
use, it is important to note that Users may not:

1. use such content for the purpose of providing other users with access on a regular or large scale
basis or as a means to circumvent access control;
2. use such content where to do so would be considered a criminal or statutory offence in any
jurisdiction, or gives rise to civil liability, or is otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association
unless explicitly agreed to by Springer Nature in writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a
systematic database of Springer Nature journal content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a
product or service that creates revenue, royalties, rent or income from our content or its inclusion as
part of a paid for service or for other commercial gain. Springer Nature journal content cannot be
used for inter-library loans and librarians may not upload Springer Nature journal content on a large
scale into their, or any other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not
obligated to publish any information or content on this website and may remove it or features or
functionality at our sole discretion, at any time with or without notice. Springer Nature may revoke
this licence to you at any time and remove access to any copies of the Springer Nature journal content
which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or
guarantees to Users, either express or implied with respect to the Springer nature journal content and
all parties disclaim and waive any implied warranties or warranties imposed by law, including
merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published
by Springer Nature that may be licensed from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a
regular basis or in any other manner not expressly permitted by these Terms, please contact Springer
Nature at

onlineservice@springernature.com

You might also like