You are on page 1of 19

Journal of Petroleum Science and Engineering 191 (2020) 107162

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

A fractional diffusion model for single-well simulation in geological media


Abiola D. Obembe
Innovation & Technology Transfer, King Fahd University of Petroleum & Minerals, Saudi Arabia

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

Keywords: In this scholarly work, a novel mathematical model for single-phase, single-well simulation in hydrocarbon
Single-well simulation reservoirs was established using the fractional calculus theory. Specifically, the existing fractional generalization
Fractional calculus of Darcy’s law is reconstructed by introducing an auxiliary parameter to ensure that the definition of the con­
Radial diffusivity equation
ventional rock permeability with dimension L2 was preserved. Subsequently, the newly constructed fractional
Finite difference method
Fractional-order derivative
flux relationship was substituted as a momentum equation in the mass conservation expression in the radial-
Grünwald–Letnikov cylindrical coordinate system. The resulting nonlocal time-radial diffusivity equation was solved numerically
using the block-centered finite-difference approximation and adopting the Grünwald–Letnikov formula for the
definition of the fractional-order derivative. The developed numerical scheme was verified by comparing the
approximate pressure solution against the semi-analytical solution of the simplified nonlocal time-radial diffu­
sivity equation and the classic radial flow diffusivity equation. Furthermore, incremental material balance checks
were performed and exhibited to ensure the conservation of mass at each time step. Finally, a detailed sensitivity
analysis was performed to evaluate the impact of the order of fractional differentiation on pressure distribution in
the reservoir and wellbore. The proposed nonlocal time-radial diffusivity equation proffers a useful model for
single-well simulation in geological media that exhibit distorted flow paths and near power-law behaviors.

1. Introduction Roeloffs, 1988) to name a few of the many citations.


For instance, Caputo (1999) postulated that the flow of particle laden
Numerical simulation of fluid flow in porous media proffers a useful fluids in porous media may obstruct the pores or trigger chemical re­
and powerful tool for effective reservoir monitoring and management of actions within the medium resulting in enlarged pores and in general
hydrocarbon bearing subterranean formations. The development of any permeability changing with time (local phenomenology). Later, Caputo
reservoir simulator comprises the following key steps; mathematical and Plastino (2003) proposed a fractional calculus model to account for
formulation, discretization (finite difference, finite element, finite vol­ the effect of the medium previously affected by the fluid during diffusion
ume, etc.), well representation (vertical wells, horizontal wells, in in porous media. Similarly, Iaffaldano et al. (2005) presented a frac­
addition to completion type, etc.), linearization, solution, and valida­ tional calculus model for diffusion of fluids in porous media to account
tion. Darcy equation was proposed in the mid of the nineteenth century for the reduction of permeability in sand layers due to the rearrange­
to proffer a phenomenological relationship between flux and potential ment of sand grains and compaction. Recently, Zhou et al. (2019) pro­
gradient (Darcy, 1856). Together with the statement of mass conserva­ posed a fractional calculus model (i.e. Swartzendruber model) to
tion, both equations constitute the classical theory and equations for the characterize non-Darcian flow in porous media and further validated the
propagation of pressure and fluids in porous media in a plethora of model with experimental data in the literature for water flows in the
existing commercial reservoir simulators. While this classical approach soil–rock mixtures. It is argued among scholars that the classical theory
may provide realistic characterization of flow in the pore throats of fails to honor the complex internal architecture and geologic structures
conventional reservoir rocks, the literature is beleaguered with experi­ which may include stratification, intercalations, and barriers that lead to
mental data and theoretical studies where the flow of fluids through distorted flow paths (Chang et al., 2019; Obembe et al., 2018c, 2017a;
rocks exhibit features which may not be represented by the classical Raghavan, 2004; Thomas et al., 2005). Other investigators demon­
theory (Bell and Nur, 1978; Brinkman, 1949; Caputo, 2000; For­ strated that flow through heterogeneous porous media fails to honor the
chheimer, 1901; Giuseppe et al., 2010; Hossain and Abu-khamsin, 2012; classical theory due to the fractal nature of naturally existing porous
Iaffaldano et al., 2005; Nigmatullin, 1986; Ochoa-Tapia et al., 2007; media (Acuna and Yortsos, 1995; Camacho-Velazquez et al., 2011;

E-mail addresses: obeabi@kfupm.edu.sa, obeabi@yahoo.com.

https://doi.org/10.1016/j.petrol.2020.107162
Received 9 August 2019; Received in revised form 10 February 2020; Accepted 4 March 2020
Available online 10 March 2020
0920-4105/© 2020 Elsevier B.V. All rights reserved.
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

Camacho Velazquez et al., 2006; Chang and Yortsos, 1990; Hardy and proposed fractional diffusion model conveniently allows for the
Beier, 1994; Razminia et al., 2015, 2014; Sahimi and Yortsos, 1990). consideration of anisotropy, stress-sensitivity, and multiphase flow etc.
Besides, the lack of a clear scale separation in unconventional reservoirs while modeling flow through porous media.
has contributed to the shortcomings of the classical theory in forecasting The plan of this paper is as follows: Section 2 details the theoretical
hydrocarbon production in shale gas reservoirs (Albinali and Ozkan, background for the re-construction of the constitutive fractional flux
2016; Chen and Raghavan, 2015; Holy and Ozkan, 2016; Obembe et al., expression and the derivation of the proposed nonlinear nonlocal time-
2017b, 2017c; Ozcan et al., 2014; Raghavan and Chen, 2018, 2016; Ren radial diffusivity equation. The implicit finite difference discretization of
et al., 2018; Ren and Guo, 2015). the proposed mathematical model including wellbore method and so­
In general, the seepage equations developed using the fractional lution methodology are presented in Section 3. The semi-analytical so­
calculus theory have been successfully applied to address the deviations lution (SAS) to the simplified version of the mathematical mode (i.e.
observed from the classical theory of fluid flow in porous media. This is uniform rock and fluid properties) is derived in Section 4. The SAS is
because such fractional diffusion models (i.e. using fractional derivative utilized later in Section 5 to verify the developed numerical model so­
operators) provide an excellent tool for the description of memory and lution (i.e. comparison of the SAS to the numerical solution). Further­
other hereditary properties of various materials and processes (Pod­ more, findings from synthetic examples of the single-well simulation are
lubny, 1998). In addition, fractional diffusion models imply the notion presented and discussed in the remainder of Section 5. The paper ends
that diffusion is irregular (i.e. anomalous diffusion) and the well re­ with some conclusions and remarks in Section 6.
sponses in such media follow a power-law behavior. A wealth of frac­
tional diffusion models have been presented in the literature using the 2. Theoretical formulation
theory of fractional calculus to exhibit the subdiffusive nature of flow in
heterogeneous porous media (Al-Rbeawi and Owayed, 2019; Amir and 2.1. The classic representation of subdiffusive flux
Sun, 2018; Awotunde et al., 2016; Chen et al., 2014; Liang et al., 2019;
Obembe et al., 2018b, 2018a; 2017d, 2017e; Razminia et al., 2019; Sun The simplest constitutive relationship describing fluid flow in porous
et al., 2009) to name a few of the numerous citations. Subdiffusion de­ media is the classical Darcy’s law expressed as
notes a transport process where the mean square displacement of the � �
βc K
diffusing particle follows a non-linear function of time (i.e. power law ⇀ ⇀
uðx ; tÞ ¼

​ rΦðx ; tÞ (1)
μo
behavior). Besides, subdiffusion implies a memory formalism where the
history of the flow process during its journey in the porous media im­
where βc ¼ 1:127 � 10 3 is a conversion factor equal, K[mD] refers to
pacts the macroscopic transport behavior. In general, subdiffusive
rock permeability and in its complete form is a tensor, μo [cp] denotes
transport has been reported to be more appropriate to describe fluid
the crude-oil viscosity, r is the gradient operator, and Φ[psia] is the
flow in a naturally fractured and disordered nano-porous media (Albi­
potential.
nali and Ozkan, 2016; Obembe et al., 2017a; Raghavan and Chen, 2019,
The potential (Φ) in Eq. (1) is related to the pressure (p) through the
2018).
expression:
Typically, these fractional diffusion models are arrived at by modi­
fying Darcy’s flux relationship via the introduction of fractional deriv­
� �
ative operators resulting in variants of non-local constitutive temporal Φ Φref ¼ p pref γ Z Zref (2)
and/or spatial flux expressions. So far, numerous analytical and nu­
merical techniques have been successfully applied to solve the resulting where the fluid gravity γ ¼ γc ρg, ρ denotes the fluid density (lbm/ft3),
fractional diffusion equations and many interesting findings have been γ c refers to the gravity conversion factor (0:21584 � 10 3), g is the ac­
reported. Some of the key differences in the literature on fractional celeration due to gravity (32:174 ft/sec 2), p is the pore pressure (psia),
diffusion models comprise of the following; the interpretation of the pref is the pressure at reference conditions, and Z is the elevation from
fractional derivative operator (e.g. Caputo, Riemann-Liouville, Grün­ datum (reference), with positive values downwards.
wald-Letnikov, etc.) and the physical interpretation of the terms asso­ In recent years, fractional constitutive laws have been shown to be
ciated with the resulting modified flux expressions. Specifically, the adequate for representing the internal architecture of rock fabric and
physical interpretation of the pseudo-permeability coefficient (Kα ) where the presence of topological, geometrical and spatial influences
associated with current expressions is ambiguous and is a major draw­ result in distorted flow paths and loss of connectivity. Specifically, for
back to the widespread adoption of these models in the oil and gas subdiffusive flow, Eq. (1) can be generalized into the fractional Darcy
industry. law given by (Caputo, 1999; Raghavan, 2011):
The goal of this study is as follows; First, to propose a fractional
diffusion model for single-phase, single-well simulation in geological � � α
media utilizing a modified fractional Darcy flux expression as the gov­ βc Kα ∂1
(3)
⇀ ⇀ ⇀
uðx ; tÞ ¼ ½rΦðx ; tÞ�
erning momentum equation in the radial-cylindrical mass conservation μo ∂t 1 α

equation. Herein, the order of fractional differentiation (α) and auxillary


where Kα [mD-day1 α] is a constant for any history-matched value of α,
parameter (σ) constitute the model parameters of the fractional diffusion 1 α
model. Second, to understand the impact of the order of fractional dif­ and ∂∂t1 α the Grünwald–Letnikov (G-L) factional derivative of order 1 α
ferentiation on reservoir and wellbore pressure response. Parametric given by the expression (MacDonald et al., 2015; Oldham and Spanier,
study on the auxiliary parameter was noted to be trivial since it is simply 1974):
a constant that multiplies a given value of permeability hence it was
neglected from further analysis. The novelty of this research is that the

2
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

∂1 α t=Δt
X ð 1Þk Γð2 αÞ relevant physics becomes problematic.
α
f ðtÞ ¼ lim Δtα 1
f ðt kΔtÞ (4) Conversely, Eq. (7) is praise-worthy in that it allows the reservoir
∂t 1 Δt→0 k!Γð2 α kÞ
k¼0
engineer to separate distinct phenomena observed in porous media.
Specifically, this flux expression clearly separates the parameters that
where Γð:Þ denotes the Gamma function and 0 < α < 1 is the order of
control sub-diffusion (i.e. α and σ) from the conventional rock perme­
fractional differentiation.
ability parameter which addresses petrophysical heterogeneity
Subdiffusive flux is characterized by 0 < α < 1 in Eq. (3) and inter­
(permeability variation with space), multiphase flow, and stress-
estingly Eq. (3) reverts to the classical Darcy’s law when α ¼ 1. Of
sensitivity etc. This praiseworthy development is illustrated through
significance, the flux expression given by Eq. (3) incorporates the
the synthetic single-well simulation examples presented in Section 5 that
memory effects of fluid transport which were neglected in Eq. (1).
account for heterogeneity, anisotropy and stress-sensitive permeability
Several scholars in petroleum engineering literature have adopted Eq.
which will be impossible to adopt using Eq. (3).
(3) as an appropriate expression for the fluid flux for subdiffusive flow
(Albinali and Ozkan, 2016; Caputo, 1999; Lemehaute and Crepy, 1983;
2.4. Derivation of nonlocal time-radial diffusivity model
Obembe et al., 2017e; Raghavan and Chen, 2019) to name a few of the
many citations. However, the main drawback to this approach is that
Fluid flow in porous media is described mathematically by
physical interpretation of Kα in Eq. (3) is not straightforward. It is argued
combining the mass conservation equation, a representative constitutive
that Kα is a dynamic property with dimension ½L2 T1 α � and consequently
equation (momentum equation), and an appropriate equation of state
is different from the conventional Darcy permeability (K) with dimen­
(Ertekin et al., 2001; Jamal et al., 2006). If we consider the
sion L2 . In addition, the literature indicates that static measurements are
two-dimensional (2D) radial-cylindrical grid geometry shown in Figure
not suitable to determine this parameter and that the only practical
(Fig. 1), the mass conservation equation is derived by acknowledging
method to estimate Kα is to match the transient pressure or flow rate
that over a given time interval (Δt ¼ t nþ1 t n ), the fluid coming from
data with an appropriate model (Ozcan, 2014). In the next sub-section,
neighboring blocks enters block (i; j) through block boundaries
the theory that constitutes the alternative approach to describe sub­ � �
diffusion in porous media is presented. (i 12;j) and (i;j 12) and leaves through its boundary blocks i þ12; j and
� �
2.2. Modified subdiffusive flux i; j þ 12 , including that fluid may enter or leave grid block ði; jÞ
through the action of an injection or a production well.
This section details the dimensionality of fractional and ordinary The statement of mass conservation for the radial-cylindrical coor­
differential operators. To begin, if the ordinary time derivative operator dinate system for 2D flow is expressed in engineering form (Jamal et al.,
is replaced by the fractional representation as follows: 2006) on page 33 as follows:
d dα
→ (5)
dt dtα
Ztnþ1 � �� Ztnþ1 � �� Ztnþ1 � ��
ur Ar �� ur Ar �� uz Az ��
Go�mez-Aguilar et al. (2012) noted the dimensional inconsistency dt dt þ dt
Bo �r Bo �r Bo � z
associated with Eq. (5) by physical inspection of the expression since the tn i 1; j tn iþ1; j tn i;j 1

(8)
2 2 2

time derivative operator dtd has a dimension of s 1, while the fractional-


Ztnþ1 � �� Ztnþ1 �� �nþ1 � �n �
order derivative operator was s . To ensure consistency with the

dtα
α
uz Az �� Vb φ φ
dt þ qsci;j dt ¼ i;j
time dimension, they introduced an auxiliary parameter (σ) to charac­ Bo �z αc Bo i;j Bo i;j
i;jþ1
terize the fractional temporal structures in the fractal space–time ge­
tn 2 tn

ometry. Accordingly, the following relationship was derived: where the porosity (φ) and oil formation volume factor (Bo ) are pressure
d 1 dα dependent, qsc is the volumetric rate at standard conditions of pressure
→ 0<α<1 (6) and temperature, Vb is the bulk volume of the grid block, αc is a volu­
dt σ1 α dtα
metric conversion factor equal to 5.615, and Ar and Az represents the
where the parameter 0 < σ < 1 has a dimension of time (day) cross-sectional area normal to the flow in r and z directions.
(Go
�mez-Aguilar et al., 2012; Go�mez et al., 2015; Kang et al., 2019). The fluid volumetric velocities in the r and z directions can be re-
Therefore, Eq. (3) can be re-written as follows: written by the algebraic expressions as follows:

� � α 0 1
βc K σ 1 α
∂1
(7)
⇀ ⇀ ⇀
uðx ; tÞ ¼ ½rΦðx ; tÞ� � �� BΦ
μo ∂t1 α
β Kr σ1 α �� ∂1 α
B i 1;j Φi;j C
C
ur jr 1 ¼ c � α B C (9)
i ; j
2 μ o ∂t1 @ Δri 1; j A
where σ and α are the phenomenological parameters that can be ob­
ri 1; j 2
2

tained by fitting the mathematical model with field data (i.e. wellbore
pressure). In addition, K is the conventional rock permeability which 0 1
may be a constant value or pressure dependent. � ��
β Kr σ1 α �� ∂ B
1 α C
BΦi;j Φiþ1;j C
ur jr 1 ¼ c � α B C (10)
iþ ; j
2 μ o riþ1; ∂t @ Δriþ12; j A
1

2.3. Significance of modified flux expression


j
2

0 1
Until now, Eq. (3) has been widely adopted to model subdiffusive
flow through porous media. Specifically, this flux representation in­ � ��
βc Kz σ1 α �� ∂1 α BΦ C
B i;j 1 Φi;j C
troduces Kα and α as the model parameters that govern the subdiffusive uz jz ¼ � α B C (11)
i;j 1
2 μ o ∂t1 @ Δzi;j 12 A
transport through porous media. However, this approach fails to
zi;j 1
2

adequately cater to other relevant physics associated with flow through


porous media e.g. multi-phase flow, anisotropy, stress-sensitivity, and
permeability impairment, etc. This challenge arises because Kα has no
straightforward interpretation and therefore acknowledging other

3
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

0 1
pressure.
� �� α BΦ C
β c Kz σ 1 α � ∂1 B i;j Φi;jþ1 C
u z jz ¼ �
� αB C (12) 3. Numerical formulation
i;jþ1
2 μo zi;jþ 1
∂t1 @ Δzi;jþ12 A
2

3.1. Development of the numerical scheme


Inspection of Eqs. (9)–(12) indicates that the mobility between two
� � � �
1 α Typically, in the finite difference method, the space-time solution’s
adjacent blocks is defined as βc Kμσ rather than the expression βμc K
o o
domain is discretized. As for single-well simulation, reservoir dis­
given in the classic Darcy flux expression. cretization involves dividing the cylinder into Nr concentric radial seg­
Substituting the expressions for the fluid volumetric velocities given ments with the well passing through the center and planes normal to the
in Eqs. (9) through (12) in Eq. (8) leads to: longitudinal axis divide the cake-like slices into Nz segments. Further­
more, rays from the center divide the radial segments into Nθ cake-like
2 3 slices (Aziz, 1979; Jamal et al., 2006). Herein, a reservoir block in a
Ztnþ1 6� �� 7 discretized reservoir is identified as block ði; jÞ where i; and j are
α
6 β c K r σ 1 α Ar �
6 � ∂1 �7 respectively the orders of the block in r, and z directions with 1< i < Nr ,
6 μ Bo Δr � Φi 1;j Φi;j 7
7dtþ
Nθ ¼ 1, 1< j < Nz . In other words, fluid flow in the θ direction is
α
n
4 o ri 1; j
∂t1 5
t
neglected, and the focus is on the two-dimensional single-well simula­
2

2 3 tion problem in r and z directions. The following habitual notations will


be defined: Δt is the timestep, Δr is the spatial mesh in the r direction, Δz
Ztnþ1 6� �� 7
6 β c K r σ 1 α Ar �
6 � ∂1 α �7 is the spatial mesh in the z direction, the coordinates of the mesh points
Φiþ1;j Φi;j 7
7dt
6 μ Bo Δr
4 o
� ∂
riþ1; j t
1 α
5 are ðri ; zj Þ and tn ¼ nΔt, and the values of the solution pðr; z; tÞ at the
center of the grid blocks are pðri ; zj ; tn Þ � pni;j � Pni;j . Note that Pni;j is the
tn
2

2 3 numerical estimate of the exact value of pðr; z; tÞ at the center of the grid
blocks ðri ; zj ; tn Þ.
Ztnþ1 6� ��
6 β c K z σ 1 α Az � ∂1 α � 7
7 (13) First, define a function ωαk such that:
6 � Φi;j Φi;j dt7
6 μ Bo Δz � ∂t1 α 1 7þ
4 o zi;j 1 5
tn ð 1Þk Γð2 αÞ
(14)
2
ωαk ¼
k!Γð2 α kÞ
2 3
MacDonald et al. (2015) showed that the coefficients ωαk can be
Ztnþ1 6� �� 7
6 β c K z σ 1 α Az �
6 � ∂1 α � 7 evaluated by means of the recursive formula:
6 μ Bo Δz � α Φi;jþ1 Φi;j dt7

4 o z i;jþ1
∂t1 5
tn 2
� �
2 α k
ωα0 ¼ 1; ωαk ¼ ωαk 1 (15)
k
Ztnþ1 �� �nþ1 � �n �
Vb φ φ
qsci;j dt ¼ i;j Therefore, the G-L fractional derivative operator defined in Eq. (4)
αc Bo i;j Bo i;j
tn may be re-written in abridged form as:
Equation (13) is the nonlocal time-radial diffusivity equation for h i
tn
single-well simulation in hydrocarbon-bearing rocks. The model is ∂1 α X
Δt

nonlinear because of the dependence of fluid and rock properties on α


f ðtÞ ¼ lim Δtα 1
ωαk ½f ðt kΔtÞ� (16)
∂t1 Δt→0
k¼0

h i
tn tn
where Δt is the integer part of Δt .
Define fluid transmissibility terms by the expression:

� �
� 1
Tri�1; j ¼ Gri�1; j σ1 α
(17)
2 2
ri�1;
2
j μo B o ri�1; j
2

� �
� 1
Tzi;j�1 ¼ Gzi;j�1; σ 1 α
(18)
2 2
zi;j�1;
2
μo Bo zi;j�1;
2

where Tri�1; j and Tzi;j�1 represent the transmissibility in r and z flow di­
2 2

rections, and Gri�1; j and Gzi;j�1 are the geometric factors between two
2 2

adjacent grid blocks in the r and z directions respectively.


Expressions for the geometric factors introduced in Eqs. (17) and
(18) for anisotropic porous media and irregular grid block distribution
were presented by Farouq Ali (1986) as follows:

Fig. 1. Discretized 2D radial-cylindrical reservoir.

4
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

β Δθ
Gri ¼ 0 1 c j 0 1 (19) ri2 ri2 1
, , (29)
1;j
2 ri2 1 ¼ � �
2 B C B C3 2
loge @ri rL 1 A loge @rL 1 ri 1 A loge ri2= 2
i
2
i
2 ri 1
6 7
6 7
6 þ 7
where r1L ¼ rw , rNL r þ1 ¼ re correspond to the inner and external boundary
4 Δzi;j Kri;j Δzi 1;j Kri 1;j 5
2 2

through which fluid may enter or leave the reservoir respectively.


Therefore, substituting Eqs. (17) and (18) in Eq. (13) leads to:
β Δθ
Griþ1;j ¼ 0
,
1 c j 0
,
1 (20)
2
2 B C B C3 2 3 2 3
loge @rL ri A loge @riþ1 rL A Ztnþ1 α Ztnþ1 α
6
iþ1
2
iþ1
2
7 6 ∂1 �7 6 ∂1 �7
6 7 4Tri 1; j 1 α Φi 1;j Φi;j 5dt þ 4Triþ1; j 1 α Φiþ1;j Φi;j 5dtþ
þ 2 ∂t 2 ∂t
6 Δzi;j Kri;j Δziþ1;j Kriþ1;j 7
4 5 tn tn

2 3 2 3
Ztnþ1 1 α Ztnþ1 α
0 1 6 ∂ �7 6 ∂1 �7
� � 4Tzi;j 1 1 α Φi;j 1 Φi;j 5dt þ 4Tzi;jþ1 1 α Φi;jþ1 Φi;j 5dtþ
2 ∂t 2 ∂t
1 2 Δθ B 2 C tn tn
2βc =
j @riþ1 ri2 1A
2 2

Gzi;j 1 ¼ " # (21) Ztnþ1 �� �nþ1


Vbi;j φ
� �n �
φ
2
Δzi;j
þ
Δzi;j 1 qsci;j dt ¼
Kzi;j Kzi;j 1
αc Bo i;j Bo i;j
tn

0 1 (30)
� �
B 2 C The bulk volume for a grid block is given by (Aziz, 1979):
2βc 1 2 Δθ ri2 1 A
@riþ1 Reservoir inner reservoir intern
=
j
2 2

Gzi;jþ1 ¼ " # (22)


2
Δzi;j Δz 0 1
Kzi;j
þ Kz i;jþ1
i;jþ1
B 2 C
Vbi;j ¼ 1 2 Δθj @riþ
=
1 ri2 1 AΔzi;j for i ¼ 1; 2; …:Nr 1 and j ¼ 1; 2; …:Nz :
2 2
Typically, for single-well simulation the fluid transmissibility terms
in the block boundaries are spaced logarithmically in r and block (31)
boundaries for bulk volume calculations are spaced logarithmically in r2
(Aziz, 1979; Ertekin et al., 2001). Accordingly, the radii for the pressure and
points, transmissibility calculations, and bulk-volume calculations are 0 1
expressed respectively as follows (Aziz, 1979): B C
Vbnr ;j ¼ 1 2 Δθj @re2
=
rn2r 1 AΔzNr ;j for i ¼ Nr and j ¼ 1; 2; …:Nz : (32)
2

� �1=
r e Nr
αlg ¼ (23) To develop the numerical scheme, the time integrals on both sides of
rw Z tnþ1
� �� Eq. (30) are approximated by the expression FðtÞdt ¼ Fnþ1 Δt ,
αlg loge αlg tn
r1 ¼ rw (24) where Δt ¼ t nþ1 t n .
αlg 1
Therefore, Eq. (30) reduces to:
For i ¼ 1; 2; …:Nr 1 � 1 α�
∂ �� � 1
∂ α � ��
riþ1 ¼ αlg ðri Þ (25) Trnþ11 1 α Φnþ1
i 1;j Φnþ1
i;j þ Trnþ11 α Φnþ1
iþ1;j Φnþ1
i;j þ
i ;j ∂t iþ ;j ∂t1
2 2

riþ1 ri � 1 α� �� � 1 α � ��
L
riþ 1 ¼ � � (26) ∂
Tznþ11 1 α Φnþ1 Φnþ1 þ Tznþ11 1

Φnþ1 Φnþ1 þ qnþ1
2 i;j 1 i;j α i;jþ1 i;j sci;j ¼
2 ∂t 2 ∂t
i;j i;jþ
loge riþ1=r
i
�� �nþ1 � �n �
Vbi;j φ φ
2
riþ1 ri2 αc Δt Bo i;j Bo i;j
2
riþ 1 ¼ � � (27)
2
2
loge riþ1 (33)
=ri 2
For slightly compressible fluids the right-hand side (RHS) of Eq. (33)
For i ¼ 2; 3; …:Nr can be represented in conservative form as follows (Jamal et al., 2006):
ri ri
(28)
1
riL 1 ¼ � � !
2 �� �nþ1 � �n � � � � �
loge ri=r Vbi;j φ φ Vbi;j φref
i 1 � ðco þ cφ Þ Pnþ1
i;j Pni;j
αc Δt Bo i;j Bo i;j αc Δt Boref
i;j

(34)

5
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

where Boref is formation volume factor (FVF) at reference conditions (i.e. Re-arranging the terms in Eq. (37) results in the following:
oil-bubble point pressure and reservoir temperature), φref is porosity at Tznþ1 Pnþ1 nþ1 nþ1
m Nr þ Trm 1;m Pm 1
reference conditions, co is the oil compressibility factor, and cφ is the
m Nr;m

h i
rock compressibility factor. Tznþ1 þ Trnþ1 þ Trnþ1 þ Tznþ1 þ Cm Pnþ1
m þ
Furthermore, for convenience, the index of the terms in Eq. (30) will
m Nr;m m 1;m mþ1;m mþNr;m

be identified with a single index ðmÞ rather than (i;j). This is achieved by Trnþ1 Pnþ1 nþ1 nþ1
Cm Pnm qnþ1
mþ1 þ TzmþNr;m PmþNr ¼
counting the grid blocks along the r direction first (horizontal sweep)
mþ1;m scm

n o n o
with i index, followed by the z direction with the j index, then the r
Tznþ1 γn Nr;m ðZm Zm Nr Þ Trnþ1 γnm 1;m ðZm Zm 1 Þ
direction again and so forth. This can be described mathematically by: m Nr;m m m 1;m

n o n o
m ¼ i þ ðj 1ÞNr (35) Trnþ1 γn ðZm Zmþ1 Þ Tznþ1 γn ðZm ZmþNr Þ
mþ1;m mþ1;m mþNr;m mþNr;m

Substituting Eq. (34) and the expression for potential difference 8 ht i 9


defined in Eq. (2) into Eq. (33) gives: >
>
>
n >
>
<X Δt h � i>
=
Tznþ1 ωαk Pnþ1
m Nr
k
Pnþ1
m
k
γ nm k
Nr;m ðZm Nr Zm Þ
� h i�
m Nr;m
>
> >
>
∂1 α
� >
: k¼1 >
;
Tznþ1 Pnþ1 Pnþ1 γ nm Nr;m ðZm Nr Zm Þ þ
m Nr;m
∂t1 α m Nr m
(38)
8 ht i 9
� 1 α h i� > n >
∂ � >
>X >
>
Trnþ1 Pnþ1 Pnþ1 γnm Zm Þ þ < Δt h i=
α m 1 m 1;m ðZm 1 �
m 1;m
∂t1 Trnþ1 ωαk Pnþ1
m 1
k
Pnþ1
m
k
γnm k
1;m ðZm 1 Zm Þ
m 1;m >
> >
>
� 1 α h i� >
: k¼1 >
;
∂ �
Trnþ1 α Pnþ1
mþ1 Pnþ1
m γnm 1;m ðZmþ1 Zm Þ þ (36)
mþ1;m
∂t1 8 ht i 9
� > n >
∂1 α h i� >
> >
� <X Δt h � i>
=
Tznþ1 Pnþ1 Pnþ1 γ nmþNr;m ðZmþNr Zm Þ þ qnþ1
scm ¼
mþNr; m
∂t1 α mþNr m Trnþ1 ωαk Pnþ1
mþ1
k
Pnþ1
m
k
γnmþ1;m
k
ðZmþ1 Zm Þ
mþ1;m >
> >
>
� � ! >
: k¼1 >
;
Vbm φref �
ðco þ cφ Þ Pnþ1
m Pnm
αc Δt Boref 8 ht i 9
m > n >
>
> >
<X Δt h i>
=
Recall the discretized expression for the G-L fractional-order deriv­ �
Tznþ1 ωαk Pnþ1 k
Pnþ1 k
γ nmþNr;m
k
ðZmþNr Zm Þ
ative given in Eq. (16) and substitute in Eq. (36) results in: mþNr;m
>
>
> k¼1
mþNr m
>
>
>
: ;
h � i
Tznþ1 Pnþ1
m Nr Pnþ1
m γnm Nr;m ðZm Nr Zm Þ þ
Of interest, for one-dimensional (1D) horizontal single-well simula­
m Nr;m

h � i tion, Eq. (38) for an interior grid block, m reduces to Eq. (39).
Trnþ1 Pnþ1
m 1 Pm nþ1
γ nm 1;m ðZm 1 Zm Þ þ
m 1;m � �
h � i Trnþ1
m 1;m
Pnþ1
m 1 Trnþ1
m 1;m
þ Trnþ1
mþ1;m
þ Cm Pnþ1 nþ1 nþ1
m þ Trmþ1;m Pmþ1 ¼
Trnþ1
mþ1;m
Pnþ1
mþ1 Pnþ1
m γ nmþ1;m ðZmþ1 Zm Þ þ 2 ht i 3
n
h � i 6X
Δt �7
6 7
Tznþ1 Pnþ1
mþNr Pnþ1
m γnmþNr;m ðZmþNr Zm Þ þ Cm Pnm qnþ1
scm Trnþ1 6 ωαk Pnþ1
m 1
k
Pnþ1
m
k
7
mþNr; m m 1;m
4 k¼1
5
8 ht i 9
> n >
>
> > 2 ht i 3
<X Δt h � i>
= n
α
Tznþ1 ωk Pnþ1
m Nr
k
Pnþ1
m
k
γ nm kNr;m ðZm Nr Zm Þ þ 6XΔt �7
6 7
> > (39)
m Nr;m
>
>
: k¼1 >
>
; Trnþ1 6 ωα Pnþ1 k
Pnþ1 k
7
mþ1;m
4 k¼1 k mþ1 m
5

8 ht i 9
> n >
>
>
<X Δt h
>
i>
= (37)

Trnþ1 ωαk Pnþ1
m 1
k
Pnþ1
m
k
γnm k
1;m ðZm 1 Zm Þ þ 3.2. Modified wellbore model
m 1;m >
> >
>
>
: k¼1 >
;
In reservoir simulation, a well is modelled as a line source/sink term.
8 ht i 9 Considering one-dimensional single-well simulation and a block-
> n >
> >
>X
< Δt h �
>
i= centered grid enclosed between the external radius r1 and the well
Trnþ1 ωαk Pnþ1 radius rw , the production rate at the wellbore can be expressed by
k
mþ1 Pnþ1
m
k
γnmþ1;m
k
ðZmþ1 Zm Þ þ
mþ1;m >
> >
>
>
: k¼1 >
; Darcy’s law for radial flow as follows:
G w1 �
8 ht i
n
9 qsc1 ¼ P1 Pwf (40)
>
> >
> μo1 Bo1
>
<X Δt h i>
=

Tznþ1 ωαk Pnþ1 k
Pnþ1 k
γnmþNr;m
k
ðZmþNr Zm Þ þ
mþNr; m
>
>
mþNr m
>
> where Pwf is the wellbore pressure and Gw1 is the well geometric factor
>
: k¼1 >
;
expressed as:
� 2π βc kr1 Δz
qnþ1 (41)
nþ1
scm ¼ Cm Pm Pnm G w1 ¼ � �
! ln rrw1
Vbm
where Cm ¼ ðco þ cφ Þ
φref
αc ðΔtÞα Boref However, for a vertical well penetrating several blocks (i.e. different
m
layers), the well production rate must be allocated among the different

6
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

layers. By extension, the contribution of wellblock m to the production h i


rate may be expressed as follows (Ertekin et al., 2001; Jamal et al., � �nþ1 X
tn
Δt
h � �i
2006): qnþ1
GWm σ 1 α
ωαk Pnþ1 k
Pnþ1 k
(50)
scm ¼ α m wfm
μom Bom Δt1 k¼0
G wm �
qscm ¼ P Pwfm (42)
μom Bom m Rearranging Eq. (50) leads to the following:
� �nþ1 h�
Furthermore, they indicated that pressure variation within a vertical GWm σ 1 α � i
wellbore is mainly attributed to the hydrostatic pressure and is given by qnþ1
scm ¼ α
Pnþ1
m Pnþ1
wfm þ δwfm (51)
μom Bom Δt 1

the expression in Eq. (43).


� where δwfm is computed from Eq. (51).
Pwfm ¼ Pwfref þ γwb Zm Zref (43)
h i
tn
where γwb ¼ γc ρBsc g, and Bo is obtained from the average wellbore pres­ Δt
Xh � �i
(52)
o

sure.
δwfm ¼ ωαk Pnþ1
m
k
Pnþ1
wfm
k

k¼1
Thus, for the case when the vertical well penetrates several blocks,
the contribution of well block m to the total well production rate can be Therefore, the wellbore pressure in wellblock m is computed from
obtained by substituting Eq. (43) into Eq. (42). Eq. (53).
� �
G wm h �i qscm μom Bom nþ1
qscm ¼ Pm Pwfref γ wb Zm Zref (44) Pnþ1 nþ1
wfm ¼ Pm þ Δt
1 α
þ δwfm (53)
μom Bom GWm σ 1 α

Given that the sum of the production rates of all wellblocks must add
up to the specified-rate-condition given by Eq. (45).
X 3.3. Solution methodology
qspc ¼ qscm (45)
mεψ w The successive Over Relaxation (SOR) iterative method was imple­
mented to handle the non-linearity of the developed nonlinear algebraic
where ψ w is the set whose elements are the blocks penetrated by the equations derived above. Specifically, for each individual grid block we
vertical well. proceed as follows:
Combining Eqs. (44) and (45), Jamal et al. (2006) derived an
expression for the flowing bottomhole pressure as follows:
I. Assume guess pressure values (Pnþ1 ) at the unknown time (i.e. n þ
ðvÞ
m
�� �h �
P Gw m
�i 1).
P P γ Z Z þ qspc
II. With Pnþ1 , estimate all fluid properties (i.e. Bo , μo etc.) using the
mεψ w μom Bom m wfref wb m ref ðvÞ

Pwfref ¼ � � (46) m
P Gw m correlations presented in Appendix B.
mεψ w μom Bom

III. Solve Eq. (38) or Eq. (39) iteratively until Pnþ1 and the obtained
ðvÞ
Besides, for the special case of equal pressure drop ðΔP ¼ Pm Pwfm Þ in m
vertically stacked wellblocks, Jamal et al. (2006) showed that the pro­ pressure (Pnþ1 ) satisfy the convergence criterion. In this study,
ðvþ1Þ
m
duction rate of each wellblock can be prorated according to the the convergence criterion is set as:
expression: �
�Pnþ1ðvþ1Þ Pnþ1ðvÞ �

� �
Gw m max � m ðvÞ
m
� � 10 6 (54)
1<m<nr � Pnþ1 �
(47)
μom Bom m
qscm ¼ � �qspsc
P Gw m
mεψ w μom Bom

IV. Once convergence is reached, the wellbore pressure at the wellblock


The simplifying assumption of equal pressure drop in perforated
(Pnþ1
wfm ) is obtained by simply solving Eq. (53).
blocks ensures a more numerically convenient treatment of the multi­
block wellbore model (i.e. avoids the implicit treatment of Pwfref in Eq.
V. The incremental material balance index, IMB , provides a useful cri­
(46) which will result in complications in the construction and solution
terion to evaluate the consistency of the pressure field obtained from
of the resulting coefficient matrix).
the numerical simulator. Ertekin et al. (2001) indicated that the IMB
Herein, the fractional calculus multiblock wellbore model is con­
should range between 0.999995 and 1.000005 at each time-step to
structed by introducing the fractional-order derivative into Eq. (40) as
ensure that the material balance is preserved. Obembe et al., 2017a,
follows (Obembe et al., 2018b):
2017b, 2018a, 2018b, presented a modified expression for the IMB for
� �
GWm σ1 α ∂1 α � subdiffusive flows as follows:
qscm ¼ Pm Pwfm (48)
μom Bom ∂t1 α �� �nþ1 � �n �
PM Vb φ φ
m¼1 Δt Bo Bo
where qscm is obtained from Eq. (47). n m m
(55)
IMB ¼ P P � nþ1 �o
Discretizing Eq. (48) for the unknown time level leads to:
M
m¼1 qnþ1
scm þ qnþ1 þ
memm lεξm qscl;m þ qnþ1
meml;m
� �nþ1 1 α � �
GWm σ1 α ∂
qnþ1
scm ¼ Pnþ1
m Pnþ1
wfm (49) where M is the total number of grid blocks, ξm is the set whose elements
μo m B o m ∂t 1 α P nþ1
are the reservoir boundaries that are shared by grid block m, ðqscm Þ
lεξm
Recall the expression for the numerical approximation for the G-L
represents the effect of the fictitious well rates introduced to account for
fractional-order derivative given in Eq. (16) and substituting into Eq.
(49) gives: flow across the reservoir boundaries, qnþ1scl;m represents the algebraic sum
of all production rates through wells within the reservoir domain, and
qnþ1
memm represents the flowrate of a fictitious well with memory in grid
block m.

7
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

4. Semi-analytical solution to the homogenous single-well where the coefficients (β; ϱ and ζ) and variables associated with Eq.
simulation model (62) are defined appropriately in Appendix A. Furthermore, the tran­
sient pressure distribution is obtained in the real time domain by uti­
In the following, the semi-analytical solution (SAS) for the nonlocal lizing the Stehfest (1970) numerical inversion algorithm. Eq. (62) will
time-radial diffusivity equation considering a homogenous one- be used in Section 5 to verify the accuracy of the numerical solution.
dimensional (1D) radial-cylindrical closed reservoir (Fig. 2) is derived
employing the Caputo definition for the time fractional-order derivative. 5. Result and discussion
Derivation of the SAS comprises the following assumptions:
In this section, three example applications are provided to illustrate
1. Homogeneous porous medium. the applications of the developed single-well simulation model. Example
2. Constant fluid and rock properties. 1 (Section 5.1) presents the single-well simulation of a homogeneous 1D
3. No-flow/closed boundary at reservoir external radius. radial-cylindrical reservoir. Model validation was established by
4. Specified production rate at the inner boundary. comparing the reservoir pressure distributions obtained from both the
SAS and the numerical model. Further model verification was performed
Mass conservation for horizontal (negligible gravity effects) fluid by comparing the pressure solutions from the developed numerical
flow in the radial coordinate system can be expressed as follows model when α ¼ σ ¼ 1 and the pressure solution obtained from the
(Muskat, 1938): classic solution of the radial diffusivity equation (i.e. based on Darcy’s
1 ∂ ∂ flux). In the second example (Section 5.2), a single-well simulation of an
ð ρur Þ ¼ ðρφÞ (56) inclined heterogeneous 1D radial-cylindrical reservoir was presented.
r ∂r ∂t
Sensitivity analysis was performed to understand the influence of key
Recall the flux relation defined in Eq. (3) and substitution in Eq. (56) model parameters (i.e. α and σ ) on pressure distribution in the reservoir
leads to (i.e. grid block pressure). Finally, single-well simulation for a 2D
� � �� �
1 ∂ βc Kr σ1r α 2πrh ∂1 α ∂p

2πhφct ∂p anisotropic and heterogenous stress-sensitive reservoir is studied in
¼ (57) Example 3 (Section 5.3). Sensitivity analysis is also conducted to
r ∂r μo Bo ∂t1 α ∂r αc Bo ∂t
investigate the effect of α on wellbore pressure only. Review of the
α 1
Introduce the operator ∂∂tα 1 , to both sides of Eq. (57) and simplifying literature indicates that for geological media exhibiting subdiffusion,
results to: 0:47 < α < 0:94 and 0 < σ � 1 (Go �mez-Aguilar et al., 2012; Go �mez
� � � � α et al., 2015; Raghavan and Chen, 2019). Moreover, in Example 2 and
1 ∂ βc Kr σ1r α ∂p φct ∂ p Example 3, the reservoir fluid is slightly compressible, and the fluid PVT
r ¼ (58)
r ∂r μo ∂r αc ∂tα properties were calculated using the correlations presented in
Appendix B. All computations in this paper were performed using
Initial condition : pðr; 0Þ ¼ pi (59) MATLAB scientific computing language.
α � �
∂1 ∂p
No flow ​ boundary ​ at ​ external ​ radius:
∂t1 α
r
∂r r¼re
¼ 0; 0 < γ<1 5.1. Example 1 model validation
(60)
Consider a 0:5 feet (ft) (0:1524 m) diameter oil well located in 60
Specified production rate at wellbore boundary condition is acres (242811 m2) spacing operating with a specified rate and the
expressed mathematically by the expression: reservoir external radius characterized as a no-flow boundary. This
� � � � synthetic radial-cylindrical reservoir is a single layer with initial pres­
2πhβc Kr σ1r α ∂1 α ∂p
qsc ¼ r ; 0<γ<1 (61) sure, thickness, horizontal permeability and porosity of 4000 psia
μo B o ∂t1 α ∂r r¼rw
(2:76 � 107 Pa), 30 ft (9:144 m), 80 mD (7:90 � 10 14 m2), and 0:23
Detailed development of the SAS of the transient pressure distribu­ respectively. The fluid in the reservoir has a density, formation volume
tion is presented in Appendix A with the resulting expression given by: factor, and viscosity of 48 lb/ft3 (768:89 kg/m3), 1:2 bbl/STB (1:2 m3/
pffiffiffiffiffiffi � pffiffiffiffiffiffi �� pi m3), and 0:45 cp (45 � 10 5 Pa-s), respectively. Other input data

b sÞ ¼ β ϱ ​ I0
Pðr; ζsα r þ K0 ζsα r þ (62) employed for computation are listed in Table 1.
s2 α s
The reservoir spatial domain is discretized into 100 grid blocks
which are spaced logarithmically according to Eqs. (24) and (25). Then
the pressure distribution in the reservoir is obtained by solving Eq. (43)
for the numerical model (NUM) and Eq. (67) for the SAS. The profiles for
reservoir pressure versus radial distance obtained from the SAS and
NUM are shown in Figs. 3 and 4 for comparison.
The figures show the curves of pressure versus radial distance when
the production time t ¼ 1 and 3 days for a well producing at a rate of 500
STB/day (Fig. 3) and 1000 STB/day (Fig. 4), respectively. The solid and

Table 1
Input parameters for Example 1.
Parameter Value

t 6 days
α 0:61
ct 3:5 � 10 6
psia 1
(5:08 � 10 8
Pa 1
)
φ 0:23
Δt 0:25 days
σr 0:8
Fig. 2. Discretized 1D radial-cylindrical reservoir.

8
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

Fig. 5. Comparisons of semi-analytical solution (SAS) and numerical solution


Fig. 3. Comparisons of semi-analytical solution (SAS) and numerical solution (NUM) to nonlocal time-radial diffusivity equation at r ¼ 0:2604 ft and 14:4897
(NUM) to nonlocal time-radial diffusivity equation at an elapsed time of (a) 1 ft for qsc equals (a) 500 STB/day, and (b) 1000 STB/day.
day and (b) 3 days (qsc ¼ 500 STB/day).

� pffiffi pffiffi pffiffi pffiffi �


b D ðrD ; sÞ ¼ 1 p
P
K0 ð s rD ÞI1 ð s reD Þ þ K1 ð s reD ÞI0 ð s rD Þ
ffiffi pffiffi pffiffi pffiffi pffiffi pffiffi (63)
s sK1 ð s ÞI1 ð s reD Þ sI1 ð s ÞK1 ð s reD Þ
The dimensionless variables in Eq. (68) in the field system of units
are defined as below:

7:08 � 10 3 Kh
PD ¼ ðpi pÞ (64)
qsc μo Bo

6:3 � 10 3 Kt
tD ¼ (65)
φμo ct rw2

r
rD ¼ (66)
rw

re
reD ¼ (67)
rw
Fig. 6 presents the pressure distribution in a closed reservoir ob­
tained from the numerical model (NUM) and the classic RDE at t ¼ 1 day

Fig. 4. Comparisons of semi-analytical solution (SAS) and numerical solution


(NUM) to nonlocal time-radial diffusivity equation at an elapsed time of (a) 1
day and (b) 3 days (qsc ¼ 1000 STB/day).

dotted lines denote the solutions from the SAS and NUM, respectively.
Examination of Figs. 3 and 4 show that there is an excellent agreement
between all the curves when t ¼ 1 day and 3 days respectively. Fig. 5
shows the pressure versus elapsed time at r ¼ 0:2604 ft (0:079 m) and
14:4897 ft (4:416 m) for qsc ¼ 500 STB/day and 1000 STB/day,
respectively.
Cleary, the plot indicates that there is an excellent overlap of the
pressure curves from the NUM and SAS.
Further verification of the developed numerical scheme was estab­
lished by comparing the solutions obtained by setting α ¼ σ ¼ 1 in the
proposed nonlocal time-radial diffusivity equation and the classic radial
flow diffusivity equation (RDE). The solution of the radial flow diffu­
sivity equation for a bounded circular reservoir with no-flow at the outer
boundary in dimensionless terms of dimensionless variables in Laplace
domain takes the form (Carslaw and Jaeger, 1940):
Fig. 6. Comparisons of classic radial diffusivity equation (RDE) and present
numerical solution (NUM) at an elapsed time of (a) 1 day and (b) 3 days
(qsc ¼ 1000 STB/day).

9
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

and 3 days respectively.


Herein, the solid and the dotted lines denote the numerical and semi-
analytical results of the reservoir pressure, respectively. The plot in­
dicates a complete overlap between both pressure curves throughout the
entire time duration which suggests that both solutions are in excellent
agreement. The pressure transient response in the closed reservoir as a
function of radial distance at qsc ¼ 1000 STB/day is depicted in Fig. 7.
The perfect agreement observed between the curves plotted in Fig. 7
establish that the classic RDE is a special case of the proposed nonlocal
time-radial diffusivity equation. In general, the numerical experiments
performed herein indicate that the developed numerical model is an
excellent prediction of reservoir pressure distribution over the range of
values of α and σ possible in a reservoir rock.

5.2. Example 2 inclined one-dimensional heterogenous reservoir

Example 2 comprises the single-well simulation of a 0:5 ft (0:1524 m)


diameter oil well located in 30 acres (121406 m2) spacing inclined at 18�
along with the formation dip, with reservoir thickness of 30 ft (9:144 m).
The radial-cylindrical reservoir is discretized into 50 grid blocks in the Fig. 8. Reservoir elevation variation with radial distance.
radial direction which are spaced logarithmically according to Eqs. (24)
and (25). The elevation depth at the center of the grid blocks along the dimensionless unit, a is the lithology factor, SVgr is the specific surface
reservoir radial direction is shown in Fig. 8. area of the grain in μm 1, and mc is the cementation exponent.
The semilogarithmic plot generated in Fig. 8 depicts a linear rela­ The realizations for the initial porosity and permeability as a func­
tionship between the elevation and radial distance, such that the tion of the radial distance represented in semi-logarithmic scales are
elevation (Z) at the wellbore (rw ) and external radius (re ) are 2971:98 ft shown in Fig. 9.
(905:8 m), and 2969:62 ft (905:14 m), respectively. Unlike the problem The input data for Example 2 including the rock and reservoir fluid
in Section 5.1, gravity effects play an important role due to the differ­ properties are documented in Table 2.
ence in elevation between adjacent grid blocks. Geological information,
completion and field data indicate that reservoir external radius is 5.2.1. Sensitivity of α on reservoir pressure
sealed to flow and the well has an open hole completion and is placed on To analyze the influence of α on the reservoir pressure, parametric
production at a rate of 1000 STB/day (1:84 � 10 3 m3/s). The reservoir analysis was conducted by independently setting α to 0:7; 0:8; and 0:9
initial porosity distribution was generated as a Gaussian field using the with other parameters kept at their default values as listed in Table 2.
MATLAB reservoir simulation toolbox (MRST) developed by SINTEF ICT The results of the incremental material balance checks performed for the
(Krogstad et al., 2015). In addition, the permeability field is estimated different values of α employed for analysis are displayed in Fig. 10.
from the porosity field adopting the Nooruddin and Hossain (2011) Clearly, Fig. 10 shows that IMB � 1 for all values of α considered
correlation: which implies that material balance is preserved at each time step. Be­
! sides, this indicates that the numerical discretization (i.e. finite differ­
1 φ2mc þ1
Kr ¼ (68) ence approximations) does not introduce significant errors to the
fg a SVgr ð1 φÞ2
2 2
numerical solution. Fig. 11 shows the effect of α on reservoir pressure
distribution at an elapsed time of 2 days and 6 days, respectively.
where Kr is permeability in μm2, fg is the shape factor in the
The semi-logarithmic plots of pressure versus radial distance in the

Fig. 7. Comparisons of classic radial diffusivity equation (RDE) and present


numerical solution (NUM) at r ¼ 0:2604 ft and 14:4897 ft Fig. 9. Randomly generated petrophysical properties variation with radial
ðqsc ¼ 1000 ​ STB =day ​ Þ. distance (a) porosity distribution and (b) permeability distribution.

10
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

Table 2 reservoir generated in Fig. 11 reveal two important features. First, it is


Base case parameters for Example 2. apparent that the magnitude of α does not impact the shape of the
Parameter Value pressure curves but rather the vertical displacement of the pressure
curves. Concretely, as α becomes smaller, the pressure curve is shifted
γAPI 37:9�
downward, indicating that the more severe the degree of subdiffusion,
Rs 200 scf/STB (23 sm3/m3)
larger pressure changes (ΔP ¼ Pi P) is observed in the reservoir.
Kw 11:8� R1/3
Second, the separation between the pressure curves increases as the
γg 0:804
simulation progresses i.e. see Fig. 11(a) and (b). The pressure history in
pi 4500 psia (3:1 � 107 Pa)
6 1 10 1
Block 1 and Block 20 is shown in Fig. 12 and some corresponding values
cs 3:5 � 10 psia (5.08 � 10 Pa )
1/3
of the block pressure at some specific time are listed in Table 3.
y N2 2:38% R �
As expected, due to production from the wellbore Fig. 12 shows that
T 120℉
there is a continuous pressure decline in both blocks with elapsed time,
0:1 days
such as when t(days) ¼ 0, 1, 2, 3, 4, 5, and 6 for α ¼ 0:7 at Block 1 in
Δt
tmax 6 days
12
Table 3, the pressure (psia) is 4500, 4042:34, 3896:50, 3784:26,
1 1:71 � 10 μm2
3687:78, 3600:74, and 3520:09, respectively. Similarly, at Block 20 in
fg a2 S2Vgr
mc 2:2 Table 3 at t(days) ¼ 0,1, 2, 3, 4, 5, and 6 for α ¼ 0:8, the pressure (psia)
σ ¼ σr 0:8 is 4500, 4241:34, 4154:60, 4082:67, 4017:32 , 3955:71, and 3896:54
respectively. Besides, it is important to note that the pressure decline in
the reservoir is more rapid as the magnitude of α becomes smaller. In
addition, Fig. 12 shows that the pressure increases with increasing the
value of α at a fixed time, such as in Block 1 in Table 3, when t ¼ 4
days, the pressure (psia) for α ¼ 0:7, 0:8; and 0:9 is 3687:78, 3779:30
and 3862:52, respectively. Similarly, in Block 20 in Table 3, when t
(days)¼ 2, the pressure (psia) for α ¼ 0:7, 0:8, and 0:9 is
4131:40, 4154:0, and 4178:02, respectively. From a reservoir manage­
ment perspective, subdiffusion implies rapid depletion of reservoir
pressure and thus initiating pressure maintenance program earlier to
ensure the reservoir is produced above bubble-point pressure (pb ).
Consequently, the value of α has a strong influence on reservoir pressure
distribution and is dependent on elapsed time.

5.3. Example 3 two-dimensional single-well simulation in a stress-


sensitive reservoir

Example 3 considers a single-well simulation in a 0:5 ft (0:1524 m)


diameter oil well located in 30 acres (121406 m2) spacing with the
reservoir top at a depth of 7000 ft (2133:6 m). The reservoir thickness
and KKHV as obtained from core data are 30 ft (9:144 m) and 0:3, respec­
Fig. 10. Incremental material balance index variation with elapsed time for
Example 2 (α ¼ 0:7,0:8, 0:9, and 1). tively. In addition, the reservoir external boundary in the radial direc­
tion is a no-flow boundary, and the well is completed in the top 20 ft
(6:096 m) only and the well production rate is specified (qspc ) at 2000
STB/day (3:68 � 10 3 m3/s). Furthermore, the reservoir bottom

Fig. 11. Pressure variation with radial distance in semilogarithmic scale at an


elapsed time of (a) 2 days and (b) 6 days.

Fig. 12. Pressure variation with elapsed time in (a) Block 1, and (b) Block 20.

11
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

Table 3
Block pressure at Block 1 and Block 20 setting σ ¼ 0:8.
Time Block 1 Block 20

(days) α ¼ 0.7 α ¼ 0.8 α ¼ 0.9 α ¼ 0.7 α ¼ 0.8 α ¼ 0.9


(psia) (psia) (psia) (psia) (psia) (psia)

0 4500 4500 4500 4500 4500 4500


1 4042.34 4053.58 4068.22 4236.37 4241.34 4248.02
2 3896.5 3942 3987.71 4131.4 4154.6 4178.02
3 3784.26 3855.17 3922.07 4045.42 4082.67 4117.99
4 3687.78 3779.3 3862.52 3968.43 4017.32 4061.96
5 3600.74 3709.72 3806.19 3896.86 3955.71 4007.99
6 3520.09 3644.23 3751.84 3828.95 3896.54 3955.26

boundary is subject to influx i.e. qz0 ¼ 1000 STB/day (1:84� 10 3 m3/


s), and the reservoir top boundary is sealed to flow (Fig. 13).
The radial-cylindrical stress-sensitive reservoir is discretized into
30 � 3 grid blocks (i.e. Nx ¼ 30 and Nz ¼ 3). The grid blocks are spaced
logarithmically according to Eqs. (24) and (25) in the radial direction
with each individual block of size 10 ft (3:048 m) in thickness. The input
initial petrophysical properties (i.e. porosity and permeability) shown in Fig. 14. Contour plot of randomly generated initial porosity distribution.
Figs. 14 and 15 respectively, are generated using the geostatistical re­
alizations from the MRST developed by SINTEF ICT (Krogstad et al.,
2015).
Specifically, the porosity field depicted in Fig. 14 was obtained using
a Gaussian field with a larger filter size in the radial direction than in the
vertical direction. Similarly, the layered realization routine in MRST was
utilized to generate a lognormally distributed horizontal permeability
field with the permeability in each layer independent of the other. The
contour plots depicted in Fig. 15 show the logarithm of the horizontal
and vertical initial permeability realizations with the horizontal
permeability field obtained by assigning mean permeability values of
100 mD, 400 mD, and 80 mD respectively, from the top to bottom layer.
Typically, permeability due to stress change influences the wellbore
response in stress sensitive reservoirs when the formation compress­
ibility is not significant. Stress-dependent permeability is modelled by
defining the permeability modulus (Nur and Yilmaz, 1985) given by the
expression:
1 ∂K
γK ¼ (69)
K ∂P
Eq. (69) leads to an exponential relationship between the grid block
permeability in and pressure as follows:
Fig. 15. Contour plot with three geological layers for (a) horizontal log
K ¼ Ki exp½ γ K ðPi PÞ� (70)
permeability realization and (b) vertical log permeability realization.
The reservoir boundary conditions and the flowing fluid (crude oil)
can be characterized by the data documented in Table 4.
The variation of the reservoir fluid PVT behavior is governed by the Table 4
Base case parameters for Example 3.
correlations provided in Appendix B. Due to the multiblock well
completion, the specified well production rate is prorated between the Parameter Value
wellblocks as expressed in Eq. (50). Table 5 lists the location of the qre 0 bbl/day (0 m3/s)
qzL 0 bbl/day (0 m3/s)
qz0 1000 bbl/day (1:84 � 10 3
m3/s)
5 9
γK 3.72�10 psia- (5:4 � 10 Pa-1)
γAPI 37:9� (870 kg/m3)
T 120 ℉
Rs 200 SCF/STB (23 sm3/m3)
γg 0:804
Pi 4000 psia (2:76 � 107 Pa)
cs 3.5 � 10 6
psia-1 (5:08 � 10 10
Pa-1)
Kw 11:8 � R1/3
y N2 2:38 mol%
Δt 0:2 days
σ r ¼ σz 0:75

Fig. 13. Discretized 2D radial-cylindrical reservoir for Example 3.

12
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

Table 5
Index of wellblocks in the reservoir.
Block location in r direction (i) Block location z direction (j) mEq. (37)

1 2 31
1 3 61

wellblocks in ði; jÞ and the corresponding ðmÞ indexes in the reservoir.

5.3.1. Sensitivity of α on wellbore pressure


A parametric study was performed to establish the impact of α on
wellbore pressure using σ ¼ σ r ¼ σz ¼ 0:95, the data listed in Tables 4
and 5 and by independently setting α ¼ 0:7,0:8, and 0:9, respectively.
The incremental balance index calculations at each time step for the
total elapsed time of 6 days is plotted in Fig. 16.
The incremental balance index versus elapsed time shown in Fig. 16
indicates that IMB � 1 for all values of α which signifies that material
balance is preserved at each time step. The influence of α on wellbore
pressure was portrayed from the curves of the wellbore pressure versus
elapsed time depicted in Fig. 17.
As expected, the wellbore pressure in Wellblock 31 and Wellblock 61 Fig. 17. Wellbore pressure variation with elapsed time in (a) Wellblock 31 and
decreases with elapsed time due to the extraction of crude oil. However, (b) Wellblock 61 (α ¼ 0:7, 0:8, 0:9, and 1).
it is seen that the wellbore pressure decline is faster as α reduces in
magnitude. Therefore, higher wellbore pressure is recorded as the
magnitude of α increases at a fixed time; as is evident from the upward Table 6
shift in the wellbore pressure curves some few days after the well was Wellbore pressure for different values of α in Wellblock 31 and Wellblock 61
placed on production. Table 6 lists some values of the wellbore pressures setting σ ¼ 0:95.
for given values of α at a specific elapsed time in Wellblock 31 and Time Wellblock 31 Wellblock 61
Wellblock 61, respectively. (days) α ¼ 0.7 α ¼ 0.8 α ¼ 0.9 α ¼ 0.7 α ¼ 0.8 α ¼ 0.9
Inspection of Table 6 points that for a given value of α; there is a
(psia) (psia) (psia) (psia) (psia) (psia)
monotonic decrease in the wellbore pressure with elapsed time, such as
when t(days)¼ 0; 1, 2, 3, 4, 5, and 6 for α ¼ 0:7 in Wellblock 31, the Pwf 0 4000 4000 4000 4000 4000 4000
1 2687.50 2714.97 2751.19 2758.49 2783.87 2817.43
(psia) is 4000, 2687:50, 2372:41, 2159:10, 1991:10, 1849:5 and 2 2372.41 2502.84 2634.36 2458.90 2581.35 2704.89
1725:67, respectively but for α ¼ 0:9 in Wellblock 31, the Pwf (psia) is 3 2159.10 2362.49 2555.74 2255.49 2446.84 2628.70
4000, 2751:19, 2634:36, 2555:74, 2491:93, 2436:19 and 2385:56, 4 1991.10 2252.27 2491.93 2094.89 2340.81 2566.52
respectively. Besides, the data indicate that at a fixed time, the smaller 5 1849.65 2159.05 2436.19 1959.21 2250.84 2511.97
6 1725.67 2076.86 2385.56 1840.03 2171.28 2462.24
the magnitude of α the lower the wellbore pressure. It is important to
note that an examination of the wellbore pressure in Wellblock 61
enumerated in Table 6 follows similar trends. Therefore, it is apparent 6. Conclusion
that the α has a strong influence on the magnitude of the wellbore
pressure. In this paper, a modified fractional diffusion model for single-well
simulation was proposed using a novel flux expression that preserves
the conventional rock permeability definition. In addition, an implicit
numerical scheme was developed to solve the resulting nonlocal time-
radial diffusivity equation adopting the G-L definition of the fractional
order derivative and block-centered finite difference approximation. Of
significance, the proposed model and formulated numerical scheme
were shown to be applicable for the single-well simulation of hetero­
geneous, anisotropic and stress-sensitive reservoir systems. Further­
more, the semi-analytical solution to the homogenous 1D radial-
cylindrical reservoir system was derived to verify the accuracy of the
approximate solution. Two synthetic examples of the single-well simu­
lation were presented to illustrate the effects of the fractional order of
differentiation (α) on pressure distribution in the grid blocks and well­
bore. It is noted that the reservoir and wellbore pressure solutions
including subdiffusion is obviously different from the classic solutions
based on Darcy’s equation. Other important conclusions are summa­
rized as follows:

1. The impact of the α on reservoir and wellbore pressure is time-


dependent with its effect more evident with the increasing of time.
This implies that the effect of the α is minimal at early times but
significant at later times. A sensitivity study performed on the α
indicated lower values of reservoir and wellbore pressure (high
Fig. 16. Incremental material balance index variation with elapsed time for
pressure changes) as the magnitude of the α decreases.
Example 3 (α ¼ 0:7,0:8, 0:9, and 1).

13
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

2. Fitting well test data with the proposed model provides the only similar works by (Holy, 2016; Ozcan, 2014; Raghavan and Chen,
realistic method of estimating the newly introduced phenomeno­ 2019) for further details.
logical parameters (i.e. α and σ ). Therefore, instead of Kh and skin
estimated from the classic radial diffusivity model from well test CRediT authorship contribution statement
data; Kh, α; and σ are all simultaneously estimated by fitting field
data or measured data with the numerical model using suitable Abiola D. Obembe: Conceptualization, Formal analysis, Methodol­
optimization algorithms. The fitting process of the proposed frac­ ogy, Validation, Visualization, Writing - original draft, Writing - review
tional diffusion model with well test data is a topic of ongoing & editing.
research by the author, although interested readers can review

Nomenclature

a Lithology factor, dimensionless


Ar Cross-sectional area of rock perpendicular to the flow of flowing fluid in r direction, ft2
Az Cross-sectional area of rock perpendicular to the flow of flowing fluid in z direction, ft2
Bo Oil formation volume factor, bbl/STB
Boref Oil formation volume factor at reference conditions, bbl/STB
co Oil compressibility, psia 1
ct Total compressibility of the system, psia 1
cφ Formation rock compressibility, psia 1
fg Shape factor, dimensionless
g ¼ 32:174 ​ Acceleration due to gravity, ft/sec2
Gr Geometric factor in r direction
GWm Well geometric factor
Gz Geometric factor in z direction
H Reservoir height, ft
i Control volume centroid counter
K Rock permeability, mD
Ki Initial permeability, mD
Kα Pseudo-permeability, mD-day1-α
Kw Watson characterization factor, � R1/3
L Reservoir length along x direction, ft
mc Cementation exponent, dimensionless
Mo Oil molecular weight, lbm/lbm mol
n Old time level
Nr Total grid blocks in r direction, dimensionless
Nθ Total grid blocks in θ direction, dimensionless
Nz Total grid blocks in z direction, dimensionless
nþ 1 New time level
P Approximate pressure, psia
pf Lasater bubblepoint pressure factor, psia
pi Initial pressure of the system, psia
Pb Bubblepoint pressure of oil without nonhydrocarbons, psia
PbN2 Bubblepoint pressure of oil with N2 present in surface gas, psia
PD Dimensionless pressure
Pref Reference pressure, psia
Pwf Flowing bottom hole pressure, psia
qsc Well rate, STB/day
qspc Specified rate at wellbore in the surface, STB/day
qz0 Specified rate at reservoir bottom boundary, bbl/day
r The radial distance along the flow direction, ft
rD Dimensionless radius
re Reservoir external radius, ft
reD Dimensionless external radius
L
ri� 1 Radial distance for transmissibility calculation, ft
2
2
ri� 1 Radial distance for bulk volume calculation, ft2
2

rw Wellbore radius, ft
Rs Solution gas oil ratio (GOR), SCF/STB
SVgr Specific surface area of the grain, μm 1
t Time, day
tD Dimensionless time
Ti Initial temperature, ℉
u Fluid flux, ft/day

14
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

ur Volumetric velocity in the r direction, ft/day


uz Volumetric velocity in the z direction, ft/day
Vb Bulk volume of grid block, ft3
xg Gas “component” mole fraction in oil
y N2 Mole fraction N2 in surface gas, %
Z Elevation from the reference point, ft
Zref Datum elevation, ft

Greek Symbols
α Order of fractional differentiation, dimensionless
αc ¼ 5:615 Volumetric conversion factor, dimensionless
αlg Defined in Eq. (23)
βc ¼ 1:127 � 10 3 Conversion factor, dimensionless
∂α
∂t α Fractional-order derivative
r Gradient operator
γ Fluid gravity, psi/ft
γ API Oil API gravity
γ c ¼ 0:21584 � 10 3 Gravity conversion factor, dimensionless
γg Gas specific gravity
γK Permeability modulus, psia 1
γo Crude-oil specific gravity
ρo Crude-oil density, lbm/ft3
ρo Crude-oil density, lbm/ft3
Δr Size of grid block in r direction, ft
Δt Timestep, day
Δz Size of grid block in z direction, ft
ΔP Reservoir pressure drop, psia
φ Porosity, fraction
φref Porosity at reference conditions, fraction
φi Initial porosity, fraction
Φ Fluid potential, psia
Φref Reference fluid potential, psia
σ Auxiliary parameter, day
γo Oil specific gravity
μo Crude-oil dynamic viscosity, cp
μob Bubblepoint oil viscosity, cp
μod Dead oil viscosity, cp
ωαk Defined in Eq. (14)

Acronyms and Field Units


API American petroleum institute
bbl Reservoir barrel
STB Standard barrel
SCF Standard cubic feet

Conversion factors
1 foot 0.3048 m
1 psia 6.894757 kPa
1cp 0.001 Pa s
1bbl/day 0.1589873 std m3/day
1 lbm/ft3 16.01846 kg/m3
1 ​ mD 0.9869233E-6 m2
1 bbl/stb 1 m3/std m3

Appendix A. Derivation of Semi-Analytical Solutions

Herein, the Laplace domain solutions for the nonlocal time-radial fractional diffusivity equation employing the Caputo definition of the fractional-
order derivative is derived in detail.
Recall Eqs. (58)–(61)
� � � � α
1 ∂ βc Kr σ1r α ∂p φct ∂ p
r ¼ (A.1)
r ∂r μo ∂r αc ∂tα

Initial Condition : pðr; 0Þ ¼ pi (A.2)

15
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

Boundary conditions:
� � � �
2πhβc Kr σ1r α ∂1 α ∂p
qsc ¼ r 0<α<1 (A.3)
μo B o ∂t1 α ∂r r¼rw

α � �
∂1 ∂p
α
r ¼0 0 < α<1 (A.4)
∂t 1 ∂r r¼re

The Laplace transform of the Caputo time fractional derivative is defined below (Ishteva, 2005):
� �
∂α
L f ðtÞ; s ¼ sα FðsÞ sα 1 f ð0Þ; 0 < α < 1: (A.5)
∂tα
Taking the Laplace Transform of Eq. (A.3) with respect to time leads to
2b � �
∂ P 1 ∂Pb � �
þ b sα 1 pðr; 0Þ
¼ ζ sα P (A.6)
∂r2 r ∂r

where ζ ¼ μo φct
αc βc Kr σ1r α .
Substitute Eq. (A.2) into Eq. (A.7) results in
2b � �
∂ P 1 ∂P b
þ b ¼ ζsα 1 pi
ζsα P (A.7)
∂r2 r ∂r
Careful observation of Eq. (A.8) indicates a second order non-homogeneous differential equation, hence the solution would consist both com­
plimentary and the particular solution.
First, the complimentary solution is of the form
pffiffiffiffiffiffi � pffiffiffiffiffiffi �
b
P ¼ A1 I0 ζsα r þ A2 K0 ζsα r (A.8)

where A1 , and A2 are coefficients to be obtained from the specified boundary conditions, I0 ð:Þ, and K0 ð:Þ are the modified Bessel functions of the first
and second kind with order zero.
Noting that the right-hand side (RHS) of Eq. (A.8) is independent of the space variable r, so P b p ¼ pi forms a particular-solution of Eq. (A.8).
s
Therefore, Eq. (A.10) is a general solution of Eq. (A.8).
pffiffiffiffiffiffi � pffiffiffiffiffiffi � pi
b ¼ A1 I0
P ζsα r þ A2 K0 ζsα r þ (A.9)
s
Applying the Laplace transform on the boundary conditions given by Eqs. (A.3)–(A.5) and imposing the initial condition leads to:
� � � �
b
∂P qsc μo Bo 1
¼ α
(A.10)
∂r r¼rw 2πrw hβc Kr σ r
1 s2 α
� �
b
∂P
¼0 (A.11)
∂r r¼re

Considering the above boundary conditions, the transient pressure distribution in the reservoir in Laplace space is governed by the expression:
� pffiffiffiffiffiffi � pffiffiffiffiffiffi �� pi
b ¼ β ϱ ​ I0
P ζsα r þ K0 ζsα r þ (A.12)
s 2 α s

where the coefficients ϱ and β are defined below


pffiffiffiffiffiffi
K1 ð ζsα re Þ
ϱ¼ pffiffiffiffiffiαffi (A.13)
I1 ð ζs re Þ
� �
qsc μo B 1
β¼ α
pffiffiffiffiffiαffi pffiffiffiffiffiffi pffiffiffiffiffiαffi (A.14)
2πrw hβc Kr σ 1r ζs ½ϱ I1 ð ζsα rw Þ K1 ð ζs rw Þ�

Appendix B. Correlations for calculating PVT properties

In this section, the correlations for calculating the PVT properties for the crude-oil are presented.

Oil gravity and molecular weight

The oil specific gravity and molecular weight are computed using:
141:5
γo ¼ (B.1)
γ API þ 131:5

16
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

� �6:58848
Kw γ0:84573
Mo ¼ o
(B.2)
4:5579

Isothermal compressibility

Calculate isothermal oil compressibility using the Frashad et al. (1996) correlation.

co ¼ 10ð Þ (B.3)
5:4531þ5:03�10 4X 3:5�10 8 X2

where
X ¼ R0:1982
s T 0:6685 γg 0:21435 γ1:0116
API P
0:1616
(B.4)

Bubblepoint pressure

The bubble point pressure is estimated with the (Lasater, 1958) correlation
pf ðT þ 459:67Þ
Pb ¼ (B.5)
γg

where the gas mole fraction in the oil and Lasater bubble factor as calculated from:
� � 1
γo
xg ¼ 1 þ (B.6)
7:521 � 10 6 Rs Mo
� �
xg 0:15649
pf ¼ exp 0:59162 (B.7)
0:33705
Finally, the Jacobson (1967) equation is adopted to correct the calculated bubblepoint pressure for the effects of nitrogen in the surface gas as
follows:
Pb N2
¼ 1:585 þ 2:86yN2 1:07 � 10 3 T (B.8)
Pb

Bubblepoint oil formation volume factor

Al-Shammasi (2001) correlation is applied to calculate the Bubblepoint oil formation volume factor.
1:410 � 10 4 Rs 4:49 � 10 4 ðT 60Þ
Bob ¼ 1 þ 5:53 � 10 7 Rs ðT 60Þ þ þ þ
γo γo
(B.9)
2:06 � 10 4 Rs γg
γo

Undersaturated Oil formation volume factor and oil density

With the obtained values of isothermal compressibility, bubblepoint oil formation volume factor and bubblepoint pressure, undersaturated oil
formation volume factor is given by the expression:
h � �i
Bo ðPÞ ¼ Bob exp co P PbN2 (B.10)

In addition, oil density is given by:


62:42796γo þ 0:0136γg Rs
ρo ¼ (B.11)
Bo

Dead oil viscosity and bubblepoint oil viscosity

Estimate the dead oil viscosity and bubblepoint oil viscosity applying the Glaso (1980) correlation and Chew and Connally (1959) correlation,
respectively.

17
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

� �
3:141 � 1010
μod ¼ 3:444
logðγAPI Þ½10:313 logðTÞ 36:447�
(B.12)
T
� �
� � 0:43þ 0:57
0:80 10ð0:00072Rs Þ
μob ¼ 0:20 þ μ (B.13)
10ð0:00081Rs Þ od

Undersaturated oil viscosity

Calculate the undersaturated oil viscosity by using the Vazquez and Beggs (1977) correlation.
� �
! 2:6 P1:187 10ð 3:9�10 5 P 5Þ
P
μo ¼ μob (B.14)
Pb N2

References Giuseppe, E. Di, Moroni, M., Caputo, M., 2010. Flux in porous media with memory :
models and experiments. Transp. Porous Media 83, 479–500. https://doi.org/
10.1007/s11242-009-9456-4.
Acuna, J.A., Yortsos, Y.C., 1995. Application of fractal geometry to the study of networks
Glaso, O., 1980. Generalized pressure-volume-temperature correlations. J. Petrol.
of fractures and their pressure transient. Water Resour. Res. 31, 527–540.
Technol. 32, 785–795. https://doi.org/10.2118/8016-PA.
Al-Rbeawi, S., Owayed, J., 2019. Analytical and numerical analysis for temporal
G�omez-Aguilar, J.F., Rosales-García, J.J., Bernal-Alvarado, J.J., C�ordova-Fraga, T.,
anomalous diffusion flow in unconventional fractured reservoirs. In: SPE Middle East
Guzm� an-Cabrera, R., 2012. Fractional mechanical oscillators. Rev. Mex. física 58,
Oil and Gas Show and Conference. Society of Petroleum Engineers.
348–352.
Al-Shammasi, A.A., 2001. A review of bubblepoint pressure and oil formation volume
G�omez, F., Morales, L., Gonz� alez, M., Alvarado, V., L�
opez, G., 2015. Fractional thermal
factor correlations. SPE Reserv. Eval. Eng. 4, 146–160. https://doi.org/10.2118/
diffusion and the heat equation. Open Phys. 13, 170–176.
71302-PA.
Hardy, H.H., Beier, R.A., 1994. Fractals in Reservoir Engineering. World Scientific.
Albinali, A., Ozkan, E., 2016. Analytical modeling of flow in highly disordered, fractured
Holy, R.W., 2016. Numerical Investigation of 1D Anomalous Diffusion in Fractured
nano-porous reservoirs. In: SPE Western Regional Meeting, 23-26 May, Anchorage,
Nanoporous Reservoirs (DISS). Colorado School of Mines. Arthur Lakes Library.
Alaska, USA. SPE-180440-MS. https://doi.org/10.2118/180440-MS.
Holy, R.W., Ozkan, E., 2016. A practical and rigorous approach for production data
Amir, S.Z., Sun, S., 2018. Physics-preserving averaging scheme based on Grunwald-
analysis in unconventional wells. In: SPE Low Perm Symposium, 5-6 May, Denver,
Letnikov formula for gas flow in fractured media. J. Petrol. Sci. Eng. (in press).
Colorado, USA. SPE-180240-MS. https://doi.org/10.2118/180240-MS.
Awotunde, A.A., Ghanam, R.A., Al-Homidan, S.S., Nasser-eddine, T., 2016. Numerical
Hossain, M.E., Abu-khamsin, S.A., 2012. Utilization of memory concept to develop heat
schemes for anomalous diffusion of single-phase fluids in porous media. Commun.
transfer dimensionless numbers for porous media undergoing thermal flooding with
Nonlinear Sci. Numer. Simul. 39, 381–395. https://doi.org/10.1016/j.
equal rock and fluid temperatures. J. Porous Media 15, 937–953. https://doi.org/
cnsns.2016.03.006.
10.1615/JPorMedia.v15.i10.50.
Aziz, K., 1979. Petroleum Reservoir Simulation. Applied Science Publishers, London.
Iaffaldano, G., Caputo, M., Martino, S., 2005. Experimental and theoretical memory
Bell, M.L., Nur, A., 1978. Strength changes due to reservoir-induced pore pressure and
diffusion of water in sand. Hydrol. Earth Syst. Sci. Discuss. 2, 1329–1357. https://
stresses and application to Lake Oroville. J. Geophys. Res. 83, 4469. https://doi.org/
doi.org/10.5194/hessd-2-1329-2005.
10.1029/JB083iB09p04469.
Ishteva, M., 2005. Properties and Applications of the Caputo Fractional Operator (Msc.
Brinkman, H.C., 1949. On the permeability of media consisting of closely packed porous
Thesis). Dept. of Math., Universit€ at Karlsruhe (TH), Sofia, Bulgaria.
particles. Appl. Sci. Res. 1, 81–86. https://doi.org/10.1007/BF02120318.
Jacobson, H.A., 1967. The effect of nitrogen on reservoir fluid saturation pressure.
Camacho-Velazquez, R., de Swaan-Oliva, A., Vasquez-Cruz, M., 2011. Interference Tests
J. Can. Pet. Technol. 6, 101–105. https://doi.org/10.2118/67-03-04.
Analysis in Fractured Formations with a Time Fractional Equation.
Jamal, H., Sm, F.A., M Rafiq, I., 2006. Petroleum Reservoir Simulation: A Basic
Camacho Velazquez, R., Fuentes-Cruz, G., Vasquez-Cruz, M.A., 2006. Decline curve
Approach. Gulf Publishing Company.
analysis of fractured reservoirs with fractal geometry. In: International Oil
Kang, J., Zhang, D., Zhou, F., Li, H., Xia, T., 2019. Numerical modeling and experimental
Conference and Exhibition in Mexico, 31 August-2 September, Cancun, Mexico. SPE-
validation of fractional heat transfer induced by gas adsorption in heterogeneous
104009-MS. https://doi.org/10.2118/104009-MS.
coal matrix. Int. J. Heat Mass Transf. 128, 492–503.
Caputo, M., 2000. Models of flux in porous media with memory. Water Resour. Res. 36,
Krogstad, S., Lie, K., Møyner, O., Møll Nilsen, H., Raynaud, X., Skaflestad, A., Ict, S.,
693–705. https://doi.org/10.1029/1999WR900299.
2015. MRST-AD – an open-source framework for rapid prototyping and evaluation of
Caputo, M., 1999. Diffusion of fluids in porous media with memory. Geothermics 28,
reservoir simulation problems:SPE-173317-MS. In: The SPE Reservoir Simulation
2113–2130. https://doi.org/10.1016/S0375-6505(98)00047-9.
Symposium. Society of Petroleum Engineers, pp. 23–25 doi:SPE-173317-MS.
Caputo, M., Plastino, W., 2003. Diffusion with space memory. In: Geodesy-The Challenge
Lasater, J.A., 1958. Bubble point pressure correlation. J. Petrol. Technol. 10, 65–67.
of the 3rd Millennium. Springer, pp. 429–435.
https://doi.org/10.2118/957-G.
Carslaw, H.S., Jaeger, J.C., 1940. Some two-dimensional problems in conduction of heat
Lemehaute, A., Crepy, G., 1983. Introduction to transfer and motion in fractal media: the
with circular symmetry. Proc. Lond. Math. Soc. 2, 361–388.
geometry of kinetics. Solid State Ionics 9 (10), 17–30. https://doi.org/10.1016/
Chang, A., Sun, H., Zhang, Y., Zheng, C., Min, F., 2019. Spatial fractional Darcy’s law to
0167-2738(83)90207-2.
quantify fluid flow in natural reservoirs. Phys. A Stat. Mech. Appl. 519, 119–126.
Liang, Y., Chen, W., Xu, W., Sun, H., 2019. Distributed order Hausdorff derivative
https://doi.org/10.1016/j.physa.2018.11.040.
diffusion model to characterize non-Fickian diffusion in porous media. Commun.
Chang, J., Yortsos, Y.C., 1990. Pressure transient analysis of fractal reservoirs. SPE Form.
Nonlinear Sci. Numer. Simulat. 70, 384–393.
Eval. 5, 31–38.
MacDonald, C.L., Bhattacharya, N., Sprouse, B.P., Silva, G.A., 2015. Efficient
Chen, C., Raghavan, R., 2015. Transient flow in a linear reservoir for space–time
computation of the Grünwald-Letnikov fractional diffusion derivative using adaptive
fractional diffusion. J. Petrol. Sci. Eng. 128, 194–202.
time step memory. J. Comput. Phys. https://doi.org/10.1016/j.jcp.2015.04.048.
Chen, S., Liu, F., Burrage, K., 2014. Numerical simulation of a new two-dimensional
Muskat, M., 1938. The flow of homogeneous fluids through porous media. Soil Sci. 46,
variable-order fractional percolation equation in non-homogeneous porous media.
169.
Comput. Math. Appl. 68, 2133–2141.
Nigmatullin, R.R., 1986. The realization of the generalized transfer equation in a medium
Chew, J.-N., Connally Jr., C.A., 1959. A viscosity correlation for gas-saturated crude oils.
with fractal geometry. Phys. Status Solidi 133, 425–430.
Trans. Am. Inst. Min. Metall. Pet. Eng. 216, 23–25.
Nooruddin, H.A., Hossain, M.E., 2011. Modified Kozeny–Carmen correlation for
Darcy, H., 1856. Les fontaines publiques de la ville de Dijon: exposition et application.
enhanced hydraulic flow unit characterization. J. Petrol. Sci. Eng. 80, 107–115.
(Google eBook), Victor Dalmont.
Nur, A., Yilmaz, O., 1985. Pore Pressure in Fronts in Fractured Rock Systems. Dep. Of
Ertekin, T., Abou-Kassem, J.H., King, G.R., 2001. Basic Applied Reservoir Simulation.
Geophysics. Stanford U., Stanford, CA.
Society of Petroleum Engineers Richardson, TX.
Obembe, A.D., Abu-khamsin, S.A., Hossain, M.E., 2018a. Anomalous effects during
Farouq Ali, S.M., 1986. Elements of Reservoir Modeling and Selected Papers.
thermal displacement in porous media under nonlocal thermal equilibrium.
Forchheimer, 1901. Wasserbewegung Durch Boden. Zeitschrift des Vereines Dtsch.
J. Porous Media 21, 161–196. https://doi.org/10.1615/JPorMedia.v21.i2.40.
Ingenieur 45, 1781–1788.
Obembe, A.D., Abu-Khamsin, S.A., Hossain, M.E., Mustapha, K., 2018b. Analysis of
Frashad, F., LeBlanc, J.L., Garber, J.D., Osorio, J.G., 1996. Empirical PVT correlations for
subdiffusion in disordered and fractured media using a Grünwald-Letnikov fractional
Colombian crude oils. In: SPE Latin America/Caribbean Petroleum Engineering
Conference. SPE-36105-MS. https://doi.org/10.2118/36105-MS.

18
A.D. Obembe Journal of Petroleum Science and Engineering 191 (2020) 107162

calculus model. Comput. Geosci. 22, 1231–1250. https://doi.org/10.1007/s10596- Raghavan, R., Chen, C., 2019. The Theis solution for subdiffusive flow in rocks. Oil Gas
018-9749-1. Sci. Technol. d’IFP Energies Nouv. 74, 6.
Obembe, A.D., Al-Yousef, H.Y., Hossain, M.E., Abu-Khamsin, S.A., 2017a. Fractional Raghavan, R., Chen, C., 2018. Time and space fractional diffusion in finite systems.
derivatives and their applications in reservoir engineering problems: a review. Transp. Porous Media 123, 173–193.
J. Petrol. Sci. Eng. 157, 312–327. https://doi.org/10.1016/j.petrol.2017.07.035. Raghavan, R., Chen, C., 2016. Rate decline, power laws, and subdiffusion in fractured
Obembe, A.D., Hasan, M., Fraim, M., 2017b. A mathematical model for transient testing rocks. In: SPE Low Perm Symposium, 5-6 May, Denver, Colorado, USA. SPE-180223-
of naturally fractured shale gas reservoirs. In: SPE Kingdom of Saudi Arabia Annual MS. https://doi.org/10.2118/180223-MS.
Technical Symposium and Exhibition, 24-27 April, Dammam, Saudi Arabia. SPE- Razminia, K., Razminia, A., Baleanu, D., 2019. Fractal-fractional modelling of partially
188058-MS. https://doi.org/10.2118/188058-MS. penetrating wells. Chaos, Solit. Fractals 119, 135–142.
Obembe, A.D., Hasan, M., Fraim, M., 2017c. An anomalous productivity model for Razminia, K., Razminia, A., Baleanu, D., 2015. Investigation of the fractional diffusion
naturally fractured shale gas reservoirs. In: SPE Kingdom of Saudi Arabia Annual equation based on generalized integral quadrature technique. Appl. Math. Model.
Technical Symposium and Exhibition. SPE-188033-MS. https://doi.org/10.2118/ 39, 86–98.
188033-MS. Razminia, K., Razminia, A., Machado, J.A.T., 2014. Analysis of diffusion process in
Obembe, A.D., Hossain, M.E., Abu-Khamsin, S.A., 2018c. Evaluation of non-fourier heat fractured reservoirs using fractional derivative approach. Commun. Nonlinear Sci.
transfer on temperature evolution in an aquifer thermal energy storage system. Numer. Simulat. 19, 3161–3170.
Transp. Porous Media 124, 825–860. https://doi.org/10.1007/s11242-018-1100-8. Ren, J., Guo, P., 2015. Anomalous diffusion performance of multiple fractured horizontal
Obembe, A.D., Hossain, M.E., Abu-Khamsin, S.A., 2017d. Variable-order derivative time wells in shale gas reservoirs. J. Nat. Gas Sci. Eng. 26, 642–651.
fractional diffusion model for heterogeneous porous media. J. Petrol. Sci. Eng. 152, Ren, J., Guo, P., Peng, S., Ma, Z., 2018. Performance of multi-stage fractured horizontal
391–405. https://doi.org/10.1016/j.petrol.2017.03.015. wells with stimulated reservoir volume in tight gas reservoirs considering anomalous
Obembe, A.D., Hossain, M.E., Mustapha, K., Abu-Khamsin, S.A., 2017e. A modified diffusion. Environ. Earth Sci. 77, 768.
memory-based mathematical model describing fluid flow in porous media. Comput. Roeloffs, E.A., 1988. Fault stability changes induced beneath a reservoir with cyclic
Math. Appl. 73, 1385–1402. https://doi.org/10.1016/j.camwa.2016.11.022. variations in water level. J. Geophys. Res. Solid Earth 93, 2107–2124.
Ochoa-Tapia, J.A., Valdes-Parada, F.J., Alvarez-Ramirez, J., 2007. A fractional-order Sahimi, M., Yortsos, Y.C., 1990. Applications of fractal geometry to porous media: a
Darcy’s law. Phys. A Stat. Mech. Appl. 374, 1–14. review. In: Annual Fall Meeting of the Society of Petroleum Engineers, New Orleans,
Oldham, K.B., Spanier, J., 1974. The Fractional Calculus: Theory and Applications of LA.
Differentiation and Integration to Arbitrary Order. Academic Press. Stehfest, H., 1970. Algorithm 368: numerical inversion of Laplace transforms [D5].
Ozcan, O., 2014. Fractional Diffusion in Naturally Fractured Unconventional Reservoirs. Commun. ACM 13, 47–49.
ProQuest Dissertations Publishing. Sun, H., Chen, W., Chen, Y., 2009. Variable-order fractional differential operators in
Ozcan, O., Sarak, H., Ozkan, E., Raghavan, R.S., 2014. A trilinear flow model for a anomalous diffusion modeling. Phys. A Stat. Mech. Appl. 388, 4586–4592.
fractured horizontal well in a fractal unconventional reservoir. In: SPE Annual Thomas, O.O., Raghavan, R.S., Dixon, T.N., 2005. Effect of scaleup and aggregation on
Technical Conference and Exhibition, 27-29 October, Amsterdam, The Netherlands. the use of well tests to identify geological properties. SPE Reserv. Eval. Eng. 8,
SPE-170971-MS. https://doi.org/10.2118/170971-MS. 248–254. https://doi.org/10.2118/77452-PA.
Podlubny, I., 1998. Fractional Differential Equations: an Introduction to Fractional Vazquez, M., Beggs, H.D., 1977. Correlations for fluid physical property prediction. In:
Derivatives, Fractional Differential Equations, to Methods of Their Solution and SPE Annual Fall Technical Conference and Exhibition. SPE-6719-MS. https://doi.
Some of Their Applications, Vol. 198 of, Mathematics in Science and Engineering. org/10.2118/6719-MS.
Academic press. Zhou, H.W., Yang, S., Zhang, S.Q., 2019. Modeling non-Darcian flow and solute transport
Raghavan, R., 2011. Fractional derivatives: application to transient flow. J. Petrol. Sci. in porous media with the Caputo–Fabrizio derivative. Appl. Math. Model. 68,
Eng. 80, 7–13. 603–615.
Raghavan, R., 2004. A review of applications to constrain pumping test responses to
improve on geological description and uncertainty. Rev. Geophys. 42.

19

You might also like