You are on page 1of 30

188 Int. J. Oil, Gas and Coal Technology, Vol. 22, No.

2, 2019

3D numerical and experimental study on upscaling


two phase relative permeability curve of naturally
fractured reservoirs

Reda Rabiee Abdel Azim*


Petroleum Engineering Department,
The American University of Kurdistan,
Duhok, Kurdistan Region of Iraq
Email: Reda.abdelrasoul@auk.edu.krd
*Corresponding author

Sheik Rahman
Petroleum Engineering School,
New South Wales University,
Sydney, Australia
Email: sheik.rahman@unsw.edu.au

Abstract: In this paper, an integrated approach of upscaling of laboratory


derived relative permeability to reservoir grid block scale under poroelastic
frame work is presented. In this approach, capillary and viscous forces are
assumed to be dominating which allows us to obtain a reasonable up-scaled
relative permeability curves. In order to improve the computation efficiency
first, the reservoir is divided into a number of grid blocks (20 m × 20 m ×
30 m) and then each block is further divided into laboratory glass bead scale
blocks (20 cm × 10 m × 2 cm). Next, relative permeability correlation
generated based on glass bead model is used to upscale the laboratory estimate
of the relative permeability curve to each fine scale grid (Fahad et al., 2017).
Levenberg-Marquardt inversion algorithm is used in the upscaling process in
order to match the oil recovery or water breakthrough time. The upscaling
approach is applied to a reservoir domain of a typical fractured basement
reservoir in Vietnam offshore. [Received: August 20, 2017; Accepted:
February 1, 2018]

Keywords: fractured reservoirs; upscaling; relative permeability curve.

Reference to this paper should be made as follows: Azim, R.R.A. and


Rahman, S. (2019) ‘3D numerical and experimental study on upscaling two
phase relative permeability curve of naturally fractured reservoirs’, Int. J. Oil,
Gas and Coal Technology, Vol. 22, No. 2, pp.188–217.

Biographical notes: Reda Rabiee Abdel Azim holds a PhD from the
University of New South Wales, Australia 2015, MSc and BSc from the
University of Cairo, Egypt in 2005 and 2010 respectively, all in Petroleum
Engineering. Prior joining Cairo University as an Assistant Professor, he
worked with Schlumberger Company (2006–2009) as a reservoir engineer for
more than three years and then with Technical Petroleum Services (TPS)
Company (2009–2011) as a Senior Consultant Petroleum Engineer. In TPS he
led research programs in the area of integrated reservoir studies and pressure
transient analysis. In 2011, he joined the University of New South Wales’s

Copyright © 2019 Inderscience Enterprises Ltd.


3D numerical and experimental study 189

School of Petroleum Engineering to do a PhD in fluid flow simulation in


fractured reservoirs. He now is a Chair of Chemical and Petroleum
Engineering Department at the American University of Ras Al Khaimah,
United Arab Emirates.

Sheik Rahman received his PhD in 1984 from Technical University of


Clausthal, Germany and gained extensive industrial experience and knowledge
by working in petroleum industry (for 12 years), and teaching (for four years at
KFUPM, Saudi Arabia and 27 years at UNSW). During the tenure of his career
he has made major contributions to hydro-thermo-mechanical and chemical
analysis of borehole stability, hydraulic stimulation of conventional and
unconventional reservoirs, multiphase flow simulation in fractured and
geothermal reservoirs and flow simulation in coal seam and shale gas
reservoirs. He has established Australia’s first drilling and well control
program.

1 Introduction

Upscaling of oil and water relative permeability curves have been performed by different
approaches including steady state methods and dynamic upscaling methods. In steady
state methods, the fluids are assumed to be in a steady state conditions, which means that
the fluid saturation distribution within the grid blocks does not change with time and
fractional flow remains constant. These assumptions are valid only over a small scale
(20 cm, or less) in which oil and water phases may come into capillary equilibrium.
There are several steady state method used in upscaling process depending on the balance
between governing forces (Kumar and Jerauld, 1996; Pickup and Stephen, 2000) which
include:
a capillary dominated flow
b viscous dominated flow
c gravity dominated flow.
In capillary dominated flow, the viscous and gravity forces are negligible and fluids are
in capillary equilibrium and saturation distribution is calculated by capillary pressure
curve. The accuracy of this method is tested by Pickup and Stephen (2000) and it has
been concluded that the method is not as accurate as dynamic two phase upscaling, but its
performance is worthy in some cases (e.g., low injection rate).
In viscous dominated flow (for example in area close to the injection well and inside
the fractures), with increase in fluid velocity, the injected fluid moves faster in the high
permeability regions (fractures). This leads to dispersion of the water flood front during
the displacement process (Dongxiao and Tchelepi, 1999). The amount of the dispersion
depends on the mobility ratio between the flowing phases. In this case, capillary forces
become smaller than the viscous forces and Buckley-Leverett formulation and Darcy’s
law can be used in upscaling process. In the case, where viscous force is dominated,
many authors used a single phase upscaling in a two phase system (Christie and Blunt
2001). In this case the dispersion of the flood front due to heterogeneities will not be
taken into account which may result in unreasonable upscaled relative permeability
curve.
190 R.R.A. Azim and S. Rahman

Gravity dominated flow depends on the fluid segregation, in which the lighter fluid
(low density) flows to high side and the high density to the low side. Since this gravity
effect depends on the density difference between fluids, gas floods are more likely to be
gravity dominated than water floods. In this case, vertical equilibrium can be assumed
and 3D simulation may be reduced to 2D.
The dynamic upscaling method is the most difficult one as it is time consuming and
requires two phase flow simulation on a fine grid scale. This method has the advantage of
taking into account of dispersion of the flood front due to heterogeneities or capillary or
gravitational forces. The dynamic upscaling method is divided into two classes:
a weighted pressure method presented by Kyte and Berry (1975)
b total mobility method presented by Stone 1991.
In weighted pressure method, the upscaled relative permeabilities are computed using
Darcy’s law and the sum of flow rates and average of pressure gradient across the grid
blocks. While, in total mobility method the upscaling works on the fractional flows. Kyte
and Berry’s (1975) method has many disadvantages that affect the upscaling results.
They include:
1 negative values of relative permeability if pressure gradient across the grid block has
the same sign as the flow rate
2 infinite relative permeability if the pressure gradient is zero
3 large error occurs in case where displacement is gravity dominated.
In addition, Kyte and Berry (1975) method requires a fine scale simulation to generate
pressure distribution in the studied domain. Fine scale simulations require high
computational resources (Barker and Thibeau, 1997). In addition, generation of fine scale
data (permeability and porosity) requires geostatistical interpolation which may create
uncertainties (Farmer, 2002). To improve the accuracy of this method and avoid
computational problems, Stone (1991) used analytical approach to compute average total
mobility and fractional flow. Darman (2000) improved the Kyte and Berry (1975)
method for cases where the gravity effects are significant. Darman’s (2000) method is
quite similar to Kyte and Berry’s (1975) method except for the transmissibility weighting
which is used to calculate the average pressure across the grid blocks.
Total mobility method are more robust than Kyte and Berry method in the sense that
infinite values can be avoided and negative values occur less frequently, but it can still
occur if the net flows of the two phase are in the opposite directions. Both methods, Kyte
and Berry and Stone’s total mobility are considered computationally expensive, but at
least Stone’s total mobility tend to be more robust than Kyte and Berry method.
The multiphase upscaling process is complicated compared to single phase and still
not well developed (Carlson, 2003; Yang et al., 2013). Existence of fractures in the
domain adds further complexity. In addition, in coarse grid, the sweep profile is
homogenous and the up-scaled relative permeability does not represent flow behaviour
appropriately. Although numerous studies have been carried out using conventional
upscaling techniques (dynamic upscaling methods) in fractured reservoir simulation
(Rossen and Shen, 1989; Talukdar et al., 2000; Ding et al., 2006) until now no
appropriate relative permeability upscaling methodology is documented using 3D flow
simulation in discrete fractures.
3D numerical and experimental study 191

In this paper, an integrated methodology is presented to upscale the relative


permeability curve from laboratory scale to reservoir grid domain scale under poroelastic
frame work. In the integrated upscaling methodology, capillary and viscous forces are
assumed to be dominating to obtain a reasonable up-scaled relative permeability curve.
The upscaling methodology is applied to a typical fractured basement reservoir domain.
Fahad et al. (2017) correlation is used to estimate the relative permeability curve for each
fine grid scale. In order to apply the concept of upscaling, two phase flow numerical
simulator under poro-elastic frame work is formulated and verified against available
analytical and numerical solution. In addition, an innovative inversion algorithm is
developed in upscaling process in order to history match the oil recovery or water
breakthrough time.
Generally, the stress has substantial effects on relative permeability measurements:
1 the measured Swi will decrease with applied stress due to water squeezing out of
small water filled pores into the larger pores filled with oil
2 Sor will increase with applied stress due to a reduction in pore volume with a constant
oil volume in the core
3 kro (Sw) will be decreased with applied stress; the effect is small but larger than the
experimental uncertainty
4 there were no significant changes in krw (Sw) with applied stress (Jones et al, 2001).
Therefore, our model incorporates the stress effects during the upscaling process.

2 Multiphase fluid flow simulation in porous fractured reservoirs

Simulation of multiphase fluid flow in naturally fractured reservoirs is a challenge due to


complexity associated with flow processes in porous matrix fracture system. A detailed
study of multiphase fluid flow for estimation of production potential of fractured
reservoirs is not well documented. In this paper, a full derivation of the multiphase fluid
flow equations in a poroelastic framework is presented to reflect the fluid flow behaviour
and evaluate fluid recovery under different driving mechanisms. A number of approaches
have been used to simulate immiscible multiphase fluid flow in fractured reservoirs.
Among these approaches most notable is the dual porosity/dual permeability approach
(Lough et al., 1996).
In this paper, we present a hybrid of single continuum and flow through discrete
fracture approach in a poroelastic frame work. Grid-based permeability tensors (3D),
which takes account of medium to small fractures, are used to simulate flow in matrix
and matrix-fracture interface and cubic law to simulate flow in discretely long fractures.
In view of this, the reservoir is divided into a number of grid blocks and short to medium
fractures, that cut these blocks, are used to calculate permeability tensors. The grid blocks
(domain) along with small and medium fractures are discretised using tetrahedral
elements in 3D domain for matrix and by triangular elements in 2D domain for fractures.
Once the block-based permeability tensors (3D) are calculated the reservoir domain along
with long fractures are discretised by tetrahedral elements for matrix as well as triangular
elements for fractures.
192 R.R.A. Azim and S. Rahman

For the purpose of multi-phase flow simulation the laboratory acquired relative
permeability is up-scaled to grid block scale. Fractured reservoirs are highly
heterogeneous in nature and therefore properties, such as the density, azimuth and dip of
fractures vary significantly from grid block to grid block which make the flow simulation
computationally exhaustive. In order to overcome this difficulty all sub-domains (grid
blocks) are grouped into a number fracture patterns based on azimuth and dip. Then the
sub-domain of each specified fracture pattern is discretised into elements of size same as
Fahad et al.’s (2017) laboratory model, 20 cm × 10 m × 2 cm. Size of these elements is
controlled using the 3D mesh generator. Then the relative permeability is defined for
each element (tetrahedral element) using Fahad et al. (2017) correlation. Then fine scale
flow simulation is carried out to up-scale the lab scale relative permeability to subdomain
scale. Once the relative permeability curve is up-scaled to each of the characteristic
subdomains, the up-scaled relative permeability are subsequently distributed to the entire
domains (1875) by matching the fracture pattern with corresponding sub-domain.
Core scale displacement experiments (both steady and unsteady state) are carried out
to derive oil and water relative permeability curves and the corresponding production
data which are then used to upscale to sub-domain reservoir scale.
Detailed derivation of the equations is presented in Abdel Azim (2016) and in
Appendix A (Abdel Azim, 2016). The poro-elastic aspects of the model, such as changes
in stresses with production draw-down or vice versa are validated against appropriate
analytical solutions and finite element tool box (Abdel Azim, 2016).

3 Governing equations for multiphase fluid flow in a poroelastic


framework

In general, behaviour of two phase fluid flow system through fractures and matrix porous
media is governed by generalised Darcy’s law and continuity equation.
The Darcy’s law is expressed as:
kij krπ
φ sπ u πs = [ − pπ + ρπ gi ] π = w, nw (1)
μπ

General continuity equation for wetting phase incorporating the concept of effective
stress can be expressed as follow:

 D  ∂ε Dc 
1− 3K  ∂t + 3K + 
 k k  ∂  ρ s  s  m
∇( pw + ρw gh)  + φ  w w  + ρw w 
ij rw m 
−∇T =  + ρwQw = 0 (2)
μ
 w wβ  ∂t β
 w  β w
  1 − φ D  ∂p 
 − 2
− 
 m ( 3Km ) 
K ∂t 

General continuity equation for non-wetting phase can be expressed as follow:

 D  ∂ε Dc 
1 − 3K  ∂t + 3K + 
 k k  ∂  ρ s  s  m
∇ ( po + ρo gh)  + φ  o o  + ρo o   + ρ Q = 0 (3)
ij ro m
−∇T = 
μ
 o oβ  ∂t β
 o  o
  1 − φ D  ∂p 
o o

 − 2
− 
 m ( 3Km ) 
K ∂t 

3D numerical and experimental study 193

where φ is the porosity of the media, Sπ is the saturation for each phase, uπs is the relative
velocity vector between fluid phase and solid phase, kij is the permeability tensor, krπ is
the relative permeability for each fluid phase π, μπ, ρπ and pπ are dynamic viscosity,
density of fluid, and fluid pressure for each phase respectively, gi is the gravity
acceleration vector, βπ is the fluid formation volume factor, Km is the bulk modulus of
solid grain, D is the elastic stiffness matrix, Qπ and represents external sources or sinks.
The computational procedure for solving fully coupled two phase fluid flow equations is
described in details in Figure 1.

Figure 1 Description of the computational procedure of two phase fluid flow calculations

Define a threshold
value of fracture
length

Starting simulation
Calculation of 3D of two phase fluid
permeability tensor for flow
short fractures using single
phase fluid flow
Initialise the reservoir model
using all rock and fluid
Mesh generation for element properties
based permeability tensor and
discrete fractures (long
fractures)
Time step
checking

Calculate nonlinear coefficient


terms

Solve unknowns Pw, Po,


displacement, and stresses on all
nodes

Increasing No
of time Solve both saturation and relative
step permeability through the relations of
t> Pc-Sw and kr-Sw
t_max

Yes
Solve the water
saturation changes If capillary
End from stabilised pressure effect is
saturation equation ignoring
194 R.R.A. Azim and S. Rahman

4 Inversion algorithm of displacement tests

The history matching of the displacement process (injection-production) is optimised by


using the Levenberg-Marquardt algorithm to minimise the error between the simulated
and experimental production data (oil and water). In this algorithm, Corey type power
law is used to create relative permeability curves during the optimisation procedures.
In step 1, a relative permeability curve is created by using Corey’s form where the
phase relative permeability is function of its saturation as described below:
b
s − s 
krw = a  w wr  (4)
 1 − swr 
d
s − s 
kro = c  o or  (5)
 1 − sor 
where krw is the water relative permeability, kro is the oil relative permeability, Swr is the
residual water saturation, Sor is the residual oil saturation, and a, b, c, and d are the
parameters, which are adjusted to achieve an acceptable history match. An initial value is
given to each of these controlling parameters (a, b, c, and d) to construct the first set of
relative permeability curves. Notable to mention, the relative permeability end points of
water and oil are determined by the experiments. Then, the simulation output (produced
volumes of oil and water with time) is used to calculate the least square function, J as
follow:
N

 (Q − Qisim )
2
J= i
obs
(6)
i =1

where Qiobs is the observed volume of produced fluid Qisim s the simulated volume of
produced fluid, and N is the total number of recovery observations to be history matched.
In step 2, the parameter ‘a’ is modified by a magnitude of ε (ε = 0.001) to estimate the
relative permeability curve which is then used to simulate two phase fluid flow to
calculate new fluid volumes (oil and water). These water and oil volumes are
subsequently used in equation (6) to calculate a new least square function, Ja and ∇Ja
then as follow:
Ja − J
∇J a = (7)
ε
In step 3, the sensitivity coefficient matrix, D is calculated as:
Q obs − Q sim
Da = (8)
ε
Now the first part of the left hand side of the Levenberg-Marquardt algorithm is obtained
and is known as Hessian matrix which is given by:
H ( X ) = DT D (9)
3D numerical and experimental study 195

In step 4, similar procedures are repeated for other controlling parameters b, c, and d until
DTD matrices are complete.
In step 5, the full formulation of left and right hand sides of Levenberg-Marquardt
algorithm are obtained. The algorithm is solved for the improvement term Δxk and the
equation used in this algorithm is as follow:

( H ( xk ) + λI ) Δxk = −∇J ( xk ) (10)

where λ is the stability factor, Δxk is used to update the control parameters (a, b, c and d)
as:
xk +1 = xk + αΔxk (11)
where α is the step size. Steps from one to five are repeated until Δxk be very small and
error between the observed and calculated volume of produced fluid is minimised.

5 Generation of the relative permeability data for core samples

The first step of the upscaling process is to simulate the unsteady state experiment so that
the history match of the production data can be performed. Core samples, in which
displacement tests are carried out, are presented in Figure 2. These core samples have a
length of 38 mm and diameter of 25.5 mm. The fracture aperture is estimated from the
volume of fluid occupying the fracture space which is 0.05 mm. The reservoir rock and
fluid properties are presented in Table 1.

Figure 2 Core samples with naturally occurring fractures from a typical basement reservoir
(see online version for colours)
196

Table 1

Experiment Length Diameter Porosity Permeability Density Viscosity Interfacial tension Capillary number
R.R.A. Azim and S. Rahman

L D ϕ k Fluids ρ µ σ Nc
Steady state
mm mm % mdarcy gm/cm3 cP mN/m vµ/σ
Corrected gas permeability 73.65 37.84 0.7 0.095
(unfractured sample)
Sample 1 (with 30º fracture) 48.8 25.4 8.28 1.226 Brine 1.063 0.95 38 10–7
Sample 2 (with 45º fracture) 25.3 25.25 16 2.135 Soltrol 0.755 1.46
Sample 3 (with 30º fracture) 30.20 25.30 10.4 1.23
Rock properties data for core samples used in relative permeability experiments
3D numerical and experimental study 197

Figure 3 shows a schematic representation of the experimental apparatus used in this


study for two phase relative permeability measurements. The apparatus consist of
injection unit, core holder, confining pressure pump and pressure recording devices. The
injection unit is a dual cylinder pump (Model 260D syringe pump) with high precision
(+0.5% of the set rate), one for brine injection and the other for oil injection. The pump
has maximum volume of 267 ml and can operate up to 52 MPa pressure. Experiment can
be run under either constant volume rate or constant pressure. The core holder is made of
stainless steel and can accommodate cylindrical core sample with dimensions from
25 mm to 60 mm in diameter and up to 100 mm in length. The confining pressure pump
is used to provide a high pressure up to 70 MPa surrounding the core sample in order to
prevent any fluid to flow through the gap between the rubber sleeve and core sample.
This pressure can be used to simulate confining pressure as well. Two pressure
transducers are used for recording and monitoring of pressure: one across the core sample
and the other at the confining pressure produced by the pump. A back pressure valve is
used to prevent the end effects. In this paper, the experiments were run under the
condition of constant injection (flow) rates.

Figure 3 A schematic representation of un-steady state experimental setup (see online version
for colours)

Prior to the displacement tests, core sample are cleaned in a soxhlet extractor using 100%
Isopropanol solvent. Next, these samples are dried in an oven at temperature of 60°C for
24 hours. Then the dry weight and dimensions (length and diameter) of the samples are
measured. The cleaned samples are placed in a vacuum pump for 12 hours to evacuate
any trapped air until the vacuum pressure is less than 0.1 m bar. De-aired brine (4% KCl)
is then allowed to flow into the sample slowly until the sample is fully submerged in
brine (about six hrs has been taken for the sample to be fully saturated). The sample is
left submerged for another eight hours to allow complete saturation. Finally both steady
and unsteady state displacement tests are carried out to measure production and their
198 R.R.A. Azim and S. Rahman

corresponding pressure. Injection of fluid is adjusted using a dual cylinder pump (Model
260D syringe pump) with high precision (+0.5% of the set rate).

Figure 4 3D generated mesh for the selected core sample

Figure 5 Water saturation profile after 1 min of water injection during the unsteady state
experiment (sample with 15° fracture) (see online version for colours)
Sw
Producer

Injector
3D numerical and experimental study 199

The relative permeability curves are generated by history matching of the production data
(the unsteady state experimental data). The history matching is carried out using the
inversion algorithm. For the purpose flow simulation in core scale 3D mesh of the core
sample is generated (see Figure 4). Matrix is represented by triangular prisms while
fracture (cutting the core sample along longitudinal axis) is represented by triangular
elements for the core sample with a single fracture cutting along the axis. The two phase
flow simulator. Figure 5 shows the water saturation profile after one minute of water
injection and Figure 6 shows the comparison between the laboratory data (produced oil
volume) and the simulation data. The relative permeability curves of the core samples
thus obtained are presented in Figures 7(a), 7(b), and 7(c).

Figure 6 Comparison between the volumes of produced oil using laboratory experiment and
simulation model (sample with 15° fracture) (see online version for colours)

The unsteady state experiment run under constant water flow rates of 0.1 cc/min for a
core sample with pore volume of 0.6 cc (see Figure 3). The pressure and cumulative oil
recovery have been recorded versus time then the history matching methodology that has
been presented in this paper used to invert the measured data to relative permeability
curve. All the three samples selected for the relative permeability include a single natural
fracture with dip angle ranging from 30° to 45°. In Figure 2 samples with varying
fracture orientation are presented. Results of these tests are presented in Figures 7(a) to
7(c). Corey type power law (Corey, 1954) is used to parameterise relative permeability
for the three core samples.
200 R.R.A. Azim and S. Rahman

Figure 7 Relative permeability curve of steady state test conducted on sample, (a) with 15°
inclination angle (b) with 45° inclination angle (c) with 30° inclination angle
(see online version for colours)

(a)

(b)
3D numerical and experimental study 201

Figure 7 Relative permeability curve of steady state test conducted on sample, (a) with 15°
inclination angle (b) with 45° inclination angle (c) with 30° inclination angle
(continued) (see online version for colours)

(c)

Residual water saturations obtained during the experiments range from 33–38% and the
oil relative permeability (end point at residual water saturation) range from 0.25 to 0.3.
Both oil and water relative permeability curves have a curvature towards higher water
saturation but near lower water saturations both curves flatten out and relative
permeability changes slowly with water saturation [see Figures 7(a) to 7(c)]. This
behaviour can be explained by the fact that there exist two flood fronts, one inside
fracture and the other inside porous matrix. The flood front sweeps faster though the
fracture than the tight matrix (basement) and as result of this the breakthrough of water
front in the fracture happens first which decreases the relative permeability rapidly with
decreasing water saturation. Once the fracture is swept through, the changes in relative
permeability are mainly due to change in saturation in matrix only. After this point, the
relative permeability of fracture-matrix system is primarily controlled by effective
permeability of the matrix.

6 Relative permeability upscaling

Core scale permeability of naturally fractured core samples cut from a typical fractured
basement reservoirs is up-scaled to a reservoir scale of size 500 m × 500 m × 90 m (see
Figure 8) [this sector has been extracted from reservoir with dimensions of (25 km ×
10 km × 300 m)] under poroelastic framework.
202 R.R.A. Azim and S. Rahman

Figure 8 3D fracture map generated using object-based model (see online version for colours)

Note: Fractures are distributed stochastically with different radius, dip and azimuth
angles using fracture intensity value of 0.1 m–1 and fractal dimension value of
(D = 1.25).
Source: Abdelazim and Rahman (2016)

6.1 Grid-based full permeability tensors


A hybrid approach is used in the simulation of fluid flow. In this approach small fractures
are considered as part of matrix and flow is simulated by using single continuum
approach. While flow through long fractures is simulated using a discrete fracture
approach. The reservoir is divided into a number of grid blocks. Grid-based full
permeability tensors for short to medium fractures are calculated using Darcy’s
diffusivity equation. Cubic law is to simulate flow through long fractures that cuts
through different grid blocks.
To calculate the effective permeability tensors which represent an average
permeability for the two structures, 3D cube is used to represent the matrix and fractures
porous media (Figure 9).
The fractured porous media is bounded in an impermeable cover with boundary
conditions for pressures (P1 and P2)
The boundary conditions are:
p ( x = 0) = p1 , p ( x = L) = p2 , J .n = 0 and v = 0 on s1

The seepage velocity calculated based on the flow rate integration over fracture surfaces
and matrix porous media and by using total volume of the block.
km
v=− ∇p (11)
μ
3D numerical and experimental study 203

where μ the fluid viscosity and p is the pressure and the continuity equation for local
seepage velocity in the matrix read as:
∇.v = 0 (12)

Figure 9 3D cube used for permeability tensor calculations


Fracture
s
Block with
volume

x=0

The hydraulic properties of fracture can be can be characterised by fracture transmissivity


(aperture) and main flow rate is set parallel and normal to fracture plane. The flow rate J
in fractures is usually defined by unit width of fracture and can be expressed by:
keff
J =− ∇s p (13)
μ

In case of the flow is parallel to fracture plane, the seepage velocity normal to the fracture
induces a pressure drop expressed by:
1
v = − ∇p (14)
μ

The effective fracture permeability of fracture can be describes by its aperture b as (in
case the fractures are empty):
b3
keff = − (15)
12
The mass conservation equation for the flow in a fracture is:
 
∇ s .J = − ( v + −v ) .n (16)
+
where n the unit vector is normal to fracture plane, v is the seepage velocity in the
matrix on the side of n and v − is the seepage velocity on the opposite side.
This transport equation is implemented with the above-mentioned boundary
conditions to calculate the permeability tensors.
204 R.R.A. Azim and S. Rahman

Therefore, the total seepage velocity over the block is obtained by integrating the
flow rates over fracture surfaces and matrix porous media. Then the results divided by the
total block volume to calculate the block effective permeability tensor.
1 − −
 −keff ∂p
vx = 
γ  γm
vx dv + 
sf
J x ds =
 μ ∂x
(17)

where sf is the surface for all fractures and γ is the matrix volume.
The details of the subsurface fracture map with detailed characteristic fracture
properties are presented in Abdel Azim et al. (2014). The input data used during the
upscaling process are presented in Table 2. The effective fracture permeability for the
whole reservoir is shown in Figure 10. The two phase flow simulator and the inversion
algorithm that presented in the above sections are used in the upscaling process. Different
steps involved in the upscaling process are presented below:
Step 1 The reservoir is divided into a number of grid blocks (1.75 no of sub-domains)
of sizes 20 m × 20 m × 30 m includes three layers in z-direction. First, four
distinct fracture patterns are identified based on azimuth and dip of fractures that
intersect the subdomains. Note also that (fractures less than the threshold value
of 40 m are considered for discrete flow simulation. These four patterns as
defined in Table 3 are selected to upscale relative permeability from laboratory
scale (core scale) to sub-domain scale. Of these characteristic fracture patters the
fracture pattern 1 is the homogeneous matrix with no intersecting fracture. The
correlation model presented by Fahad et al. (2017) based on a laboratory glass
bead model with dimensions 20 cm × 10 cm × 2 cm, is used to up-scale the core
scale relative permeability to the subdomain scale (20 m ×
20 m × 30 m).
The relative permeability curve for fracture pattern 1 (homogenous) is obtained
by acquiring displacement data for core samples from the basement reservoir
with homogeneous matrix.
Fahad et al. (2017) conducted enormous number of experiments using glass
bead model and based on the results of these experiments, the authors developed
a correlation for relative permeability estimation of fractures porous system. The
correlation as follows:

kro = 1 + ( −0.015 × n2f + 0.155 × n f ) + ( 0.16 × e−4.01(1−sin ψ ) ) − 0.085 × sin θ  × kref (18)

nw−ref
 S − S wr 
krw = krw− max  w  (19)
 1 − Sw 

Swr = 1 + ( −0.059 × n2f + 0.355 × n f ) + ( 0.304 × e−3.7(1−sin ψ ) ) − 0.19 × sin θ  × Sref (20)

where kro is the oil relative permeability, kref is the reference oil relative
permeability which is kro of the homogenous glass bead pack (Fahad et al.,
2017), nw–ref and Sref are Corey exponent and residual saturation for glass bead
pack respectively, nf is the number of fractures in the system, θ is the fractures
orientation towards the flow direction, ψ and is the angle between the fracture
3D numerical and experimental study 205

with a reference fracture (one fracture exist in the system is assumed to be a


reference).
Results of this correlation compared with the relative permeability obtained by
simulation studies are presented in Figure 11. It can be seen from Figure 11 that
the simulated oil relative permeabilities and that obtained based on the derived
correlation and simulation are in a good agreement. From the comparison, it is
clear that the correlation has the power to obtain the relative permeability of
fracture systems (different fractures orientations) but this correlation is limited
to the glass bead size. However, the correlation will be used in upscaling
process. The obtained relative permeability curve for fracture pattern (1) is
presented in Figure 11.
Step 2 The three characteristic subdomains each of which is intersected by one of the
three fracture patterns, are discretised into (20 cm × 10 m × 2 cm) element size.
This element size is controlled using the 3D mesh generator. Each element size
represents Fahad et al. (2017) glass bead model (laboratory scale) (see Figures
13, 14). Then the relative permeability is defined for each element (tetrahedral)
using Fahad et al. (2017) correlation. The rock and fluid properties include
residual water saturation and relative permeability end points for oil and water
phase are taken from the laboratory experiments carried out on a core samples
with fracture.

Figure 10 3D fracture equivalent permeability the studied reservoir (see online version
for colours)

k ,md
Kxx, mD

Selected Sector

z
y
x
206 R.R.A. Azim and S. Rahman

Step 3 Next, the fine scale numerical simulation run is carried out using the developed
multiphase numerical simulator at constant pressure drop between the injector
and the producer. The oil and water production rates are calculated with time at
the producer for these four characteristic fracture patterns (sub-domains). These
four production data are referred to as original production data. The inversion
technique is then used to match the calculated oil recovery or the water
breakthrough time (time at which water break in to the wellbore) to produce
relative permeability curve that represents the fluid flow behaviour within the
subdomains.

Figure 11 Estimated relative permeability curve for using (a) Fahad et al. (2017) correlation and
(b) simulation work presented in this paper (see online version for colours)

Table 2 Reservoir inputs data for a typical fractured basement reservoir

Parameter Value
Reservoir dimensions 500 m × 500 m × 90 m
Matrix permeability 0.1 mD
Matrix porosity 2%
Fracture aperture 7.06 × 10–3 mm
Initial fracture intensity 0.15 m–1
Fractal dimension (D) 1.25
Initial reservoir pressure 5,063 psia (34.9 MPa)
Injection pressure ( injection case) 6,409 psia (44.2 MPa)
Fluid viscosity 1.38 cp
Fluid compressibility 10–5 psi–1
Source: Farag et al. (2010)
3D numerical and experimental study 207

Step 4 Once the relative permeability curve is up-scaled to each of the four
characteristic subdomains as described in step 2, the up-scaled relative
permeability are subsequently distributed to the entire domain comprising of
1,875 subdomains by matching of the fracture pattern with the corresponding
sub-domain. Noteworthy, so far we have now obtained four characteristic
production data and their corresponding relative permeability which describes
two phase flow behaviour in 1,875 sub-domains. This was done in order to save
extensive computation time.

Figure 12 (a) Fracture map (fractures larger than 40 m) from a cross section of the basement
reservoir and (b) fracture map is divided into 625 sub-domains in 2D and 1,825
sub-domains in 3D (see online version for colours)

(a) (b)

Figure 13 Fracture map is divided into 625 sub-domains in 2D and 1,825 sub-domains in 3D
(left) (see online version for colours)

Note: Further, each sub-domain (20 m × 20 m × 30 m) is further divided into lab glass
bead model domain of size 20 cm ×10 cm × 2 cm (right).
208 R.R.A. Azim and S. Rahman

Figure 14 Estimated relative permeability curve for fracture pattern (1) obtained by using
unfractured (homogenous) core sample (see online version for colours)

Table 3 Four dominated fractured patterns of size 20 m × 20 m × 30 m

Fracture Fracture Fracture Fracture


Multi-fracture system
pattern (1) pattern (2) pattern (3) pattern (4)
Average dip angle (β) Homogenous 82° ≥ β > 70° 88° ≥ β > 82 β <= 90°
and azimuth angle 50° ≥ α > 40° 115° ≥ α > 90° 175° ≥ α >
(α) of fracture system 150°

7 Results and discussion

The relative permeability curve for fracture pattern 1 (homogenous) is obtained based on
homogeneous sample.
The results of the water saturation distribution after 3 hrs of water injection are shown
in Figures 15, 16, and 17. These figures show that water displaces oil in the discrete
fractures first and then the oil in the matrix porous media depending on the permeability
tensor value. The three characteristic oil and water production rates for fractured patterns
are shown in Figures 18, 19, 21, 22, 24, and respectively. From these figures it can be
seen that the water breakthrough takes place after 6 hrs, 3 hrs, and 1 hr of water injection
for fracture pattern 2, 3, and 4 respectively and this breakthrough time estimated by the
upscaling procedure match well with that estimated by using Fahad et al.’s (2017)
correlation. The characteristic water production rate curves between breakthroughs till
end of simulation time match reasonably well with that estimated by Fahad et al. (2017)
correlation. The up-scaled relative permeability curves for the different fracture patterns
(see Table 1) are shown in Figures 20, 23, and 26.
3D numerical and experimental study 209

Figure 15 Water saturation profile for fracture pattern (2) after 0.5 hr of water injection at
Pinitial = 34.9 MPa, Pinj = 44.2 MPa, ∆P = 3.1 MPa, σH = 33.1 MPa, σh = 33.1 MPa,
and σv = 41.3 MPa (see online version for colours)
SW

Note: Block size (20 m × 20 m × 30 m).

Figure 16 Water saturation profile for fracture pattern (3) after 0.5 hr of water injection at
Pinitial = 34.9 MPa, Pinj = 44.2 MPa, ∆P = 3.1 MPa, σH = 33.1 MPa, σh = 33.1 MPa,
and σv = 41.3 MPa (see online version for colours)

SW

Note: Block size (20 m × 20 m × 30 m).


Capillary pressure in discrete fracture tends to be negligible due to large hydraulic radius
and by reducing the number of fracture or fracture intensity in the studied domain, the
capillary force increases and the upscaled relative permeability is shifted towards the
matrix relative permeability (see Figure 20)
210 R.R.A. Azim and S. Rahman

Figure 17 Water saturation profile for fracture pattern (4) after 0.5 hr of water injection at
Pinitial = 34.9 MPa, Pinj = 44.2 MPa, ∆P = 3.1 MPa, σH = 33.1 MPa, σh = 33.1 MPa,
and σv = 41.3 MPa (see online version for colours)

SW

Note: Block size (20 m × 20 m × 30 m).

Figure 18 Comparison between original and upscaled oil production rate for fracture pattern (2)
(see Table 3) with Pinitial = 34.9 MPa, Pinj = 44.2 MPa, ∆P = 3.1 MPa, σH = 33.1 MPa,
σh = 33.1 MPa, and σv = 41.3 MPa (see online version for colours)

Note: Block size (20 m × 20 m × 30 m).


Source: Estimated based on Fahad et al.’s (2017) correlation
It is notable to mention that, the oil relative permeability and the intersection point shifts
upward if the fracture intensity (number of fractures) is increased (see Figure 23).
Moreover, the water relative permeability has no constant trend and it moves downward
and upward with increasing the fracture intensity.
3D numerical and experimental study 211

Figure 19 Comparison between original and upscaled water production rate for fracture pattern
(2) (see Table 3) with Pinitial = 34.9 MPa, Pinj = 44.2 MPa, ∆P = 3.1 MPa, σH = 33.1
MPa, σh = 33.1 MPa, and σv = 41.3 MPa (see online version for colours)

Note: Block size (20 m × 20 m × 30 m).


Source: Estimated based on Fahad et al.’s (2017) correlation

Figure 20 Upscaled relative permeability curves for fracture pattern (2) in a block size (20 m ×
20 m × 30 m) with average fractures dip angle of 76° (see online version for colours)
212 R.R.A. Azim and S. Rahman

Figure 21 Comparison between original and upscaled oil production rate for fracture pattern (3)
(see Table 3) with Pinitial = 34.9 MPa, Pinj = 44.2 MPa, ∆P = 3.1 MPa, σH = 33.1 MPa,
σh = 33.1 MPa, and σv = 41.3 MPa (see online version for colours)

Note: Block size (20 m × 20 m × 30 m).


Source: Estimated based on Fahad et al.’s (2017) correlation

Figure 22 Comparison between original and upscaled water production rate for fracture pattern
(3) (see Table 3) with Pinitial = 34.9 MPa, Pinj = 44.2 MPa, ∆P = 3.1 MPa, σH = 33.1
MPa, σh = 33.1 MPa, and σv = 41.3 MPa (see online version for colours)

Note: Block size (20 m × 20 m × 30 m).


Source: Estimated based on Fahad et al.’s (2017) correlation
3D numerical and experimental study 213

Figure 23 Upscaled relative permeability curves for fracture pattern (3) in a block size
(20 m × 20 m × 30 m) with average fractures dip angle of 89.5° (see online version
for colours)

Figure 24 Comparison between original and upscaled oil production rate for fracture pattern (4)
(see Table 3) with Pinitial = 34.9 MPa, Pinj = 44.2 MPa, ∆P = 3.1 MPa, σH = 33.1 MPa,
σh = 33.1 MPa, and σv = 41.3 MPa (see online version for colours)

Note: Block size (20 m × 20 m × 30 m).


Source: Estimated based on Fahad et al.’s (2017) correlation
214 R.R.A. Azim and S. Rahman

Figure 25 Comparison between and upscaled water production rate for fracture pattern (4) (see
Table 3) with Pinitial = 34.9 MPa, Pinj = 44.2 MPa, ∆P = 3.1 MPa, σH = 33.1 MPa, σh =
33.1 MPa, and σv = 41.3 MPa (see online version for colours)

Note: Block size (20 m × 20 m × 30 m).


Source: Estimated based on Fahad et al.’s (2017) correlation
Figure 26 Upscaled relative permeability curves for fracture pattern (4) in a block size
(20 m × 20 m × 30 m) with average fractures dip angle of 90° (see online
version for colours)
3D numerical and experimental study 215

The calculated directional effective permeability tensors using periodic boundary


conditions for small to medium fractures influenced the movement direction to oil and
water relative permeability.
As viscous forces increase in case we have high permeability region or greater
pressure difference, the water relative permeability shifts upwards (see Figure 26), that
means the dynamic behaviour of relative permeability is function of reservoir facies and
hysteresis not only with respect to drainage or imbibition but also flow rates. Therefore,
the dynamic behaviour of relative permeability must take into account the daily
production rates increases or decreases.

8 Conclusions

In this paper, an integrated upscaling technique is presented to upscale the relative


permeability curve from laboratory scale to grid block scale for a naturally fractured
basement reservoir. The upscaling was performed by dividing the reservoir into a number
of grid blocks. First the 3D effective permeability tensors for each block were estimated
by simulating multiphase fluid flow in discrete fractures. During the simulation process
we coupled the optimised relative permeability and Fahad et al. correlation model which
is a function of number of fractures, fracture aperture and fracture orientation. The
optimised relative permeability curves, thus obtained are used to simulate flow in glass
bead model (20 cm × 10 m × 2 cm) with single and multiple fractures of different
orientations to validate the numerical procedure and check if the core scale relative
permeability comply with correlation model derived by Fahad et al. (2017). The relative
permeability curves are then up scaled to the reservoir grid scale in multiple steps in
order to retain high level of accuracy.
In general, the presented technique eliminate the use of dual porosity model in which
it does not consider fluid distribution within the matrix blocks during the simulation
period and only can be applied to small number of large scale interconnected fractures.
The results of this study show that the upscaled relative permeability process is
significantly influenced by fracture orientation and connectivity and follow dynamic
behaviour depending on the driving forces, direction of flow and fracture properties. In
addition, simulation of fluid flow using discrete fracture model provides new insights into
the displacement process of naturally fractured reservoirs.
Fractures in this study are considered continuous from reservoir subdomain to
sub-domain where fractures intersect the boundaries of the matrix. 3D features account
for flow behaviour, such as gravity.
The history matching of oil production and water breakthrough time that was
generated by using small grid blocks is optimised to be up-scaled by using the
Levenberg-Marquardt algorithm to minimise the error between the simulated and
experimental production data. In this algorithm, Corey’s correlation is used to create
different relative permeability curves during the optimisation procedures. Minimum error
in produced volumes was set as the objective function of the algorithm. The objective
function is minimised by using a certain convergence criterion. The forward modelling is
a 3D multiphase fluid flow simulator based on flow through discrete fracture approach.
216 R.R.A. Azim and S. Rahman

References
Abdel Azim, R.G.D.N., Rahman, S.S., Tyson, S. and Regenauer-Lieb, K. (2014)
‘3D poro-thermo-elastic numerical model for analysing pressure transient response to improve
the characterization of naturally fractured geothermal reservoirs’, Geothermal Resources
Council, Vol. 38, No. 1, pp.907–915.
Abdel Azim, R. (2016) ‘An integrated approach for relative permeability estimation of fractured
porous media: laboratory and numerical simulation studies’, Journal of Petroleum Exploration
and Production Technology, pp.1–18.
Abdel Azim, R. and Rahman, S.S. (2016) ‘Estimation of permeability of naturally fractured
reservoirs by pressure transient analysis: an innovative reservoir characterisation and flow
simulation’, Journal of Petroleum Science and Engineering, September, Vol. 145,
pp.404–422.
Barker, J.W. and Thibeau, S. (1997) ‘A critical review of the use of pseudorelative permeabilities
for upscaling’, SPE Reservoir Engineering, Vol. 12, No. 2, pp.138–143.
Carlson, M. (2003) Practical Reservoir Simulation: Using, Assessing, and Developing Results,
PennWell Books, Houston, TX 77027, USA.
Christie, M. and Blunt, M. (2001) ‘Tenth SPE comparative solution project: a comparison of
upscaling techniques’, SPE Reservoir Evaluation & Engineering, Vol. 4, No. 4, pp.308–317.
Corey, A.T. (1954) ‘The interrelation between gas and oil relative permeabilities’, Producers
Monthly, Vol. 19, No. 1, pp.38–41.
Darman, N.B.H. (2000) Upscaling of Two-Phase Flow in Oil-Gas Systems, PhD thesis, Heriot-Watt
University.
Ding, Y., Basquet, R. and Bourbiaux, B. (2006) ‘Upscaling fracture networks for simulation of
horizontal wells using a dual-porosity reservoir simulator’, SPE Reservoir Evaluation &
Engineering, Vol. 9, No. 5, pp.513–520.
Dongxiao, Z. and Tchelepi, H. (1999) ‘Stochastic analysis of immiscible two-phase flow in
heterogeneous media’, SPE Journal, Vol. 4, No. 4, pp.380–388.
Fahad, M. (2013) Simulation of Fluid Flow and Estimation of Production from Naturally Fractured
Reservoirs, PhD thesis.
Fahad, M., Hussain, F., Rahman, S.S. and Cinar, Y. (2017) ‘Experimental investigation of
upscaling relative permeability for two phase flow in fractured porous media’, Journal of
Petroleum Science and Engineering, January, Vol. 149, pp.367–382.
Farag, S.M., Mas, C., Maizeret, P-D., Li, B. and Le, H.V. (2010) ‘An integrated workflow for
granitic basement reservoir evaluation’, SPE Reservoir Evaluation & Engineering, Vol. 13,
No. 6, pp.893–905.
Farmer, C. (2002) ‘Upscaling: a review’, International Journal for Numerical Methods in Fluids,
Vol. 40, Nos. 1–2, pp.63–78.
Jones, C., Al-Quraishi, A.A., Somerville, J.M. and Hamilton, S.A. (2001) Stress Sensitivity of
Saturation and End-Point Relative Permeabilities, Society of core analysts, Edinburgh,
Scotland.
Kumar, A.T. and Jerauld, G. (1996) ‘Impacts of scale-up on fluid flow from plug to gridblock scale
in reservoir rock’, Paper presented at the SPE/DOE Improved Oil Recovery Symposium.
Kyte, J.R. and Berry, D. (1975) ‘New pseudo functions to control numerical dispersion’, Society of
Petroleum Engineers Journal, Vol. 15, No. 4, pp.269–276.
Lough, M.F., Lee, S.H. and Kamath, J. (1996) ‘A new method to calculate the effective
permeability of grid blocks used in the simulation of naturally fractured reservoirs’, Paper
presented at the SPE Annual Technical Conference.
Pickup, G.E. and Stephen, K.D. (2000) ‘An assessment of steady-state scale-up for small-scale
geological models’, Petroleum Geoscience, Vol. 6, No. 3, pp.203–210.
Rossen, R. and Shen, E. (1989) ‘Simulation of gas/oil drainage and water/oil imbibition in naturally
fractured reservoirs’, SPE Reservoir Engineering, Vol. 4, No. 4, pp.464–470.
3D numerical and experimental study 217

Stone, H. (1991) ‘Rigorous black oil pseudo functions’, Paper presented at the SPE Symposium on
Reservoir Simulation.
Talukdar, M., Banu, H., Torsæter, O. and Kleppe, J. (2000) ‘Applicability and rate sensitivity of
several up scaling techniques in fractured reservoir simulation’, Paper presented at the SPE
International Petroleum Conference and Exhibition, Mexico.
Yang, Y., Wang, X., Wu, X-H. and Bi, L. (2013) ‘Multiphase upscaling using approximation
techniques’, Paper presented at the SPE Reservoir Simulation Symposium.

You might also like