You are on page 1of 23

Journal of Petroleum Science and Engineering 189 (2020) 106956

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Inspectional and dimensional analyses for scaling of low salinity


waterflooding (LSWF): From core to field scale
Sabber Khandoozi , Mohammad Reza Malayeri , Masoud Riazi *, Mojtaba Ghaedi
Enhanced Oil Recovery (EOR) Research Centre, IOR/EOR Research Institute, Shiraz University, Shiraz, Iran

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

Keywords: In the past decade, Low Salinity Waterflooding (LSWF) has gained increased attention as an Enhanced Oil Re­
Dimensional analysis covery (EOR) approach. Given a large number of influential parameters, the present study aimed at finding
Inspectional analysis independent scaling groups for LSWF using both Dimensional (DA) and Inspectional Analyses (IA). Scaling was
Low salinity waterflooding
performed using a 2-D vertical reservoir model with homogeneous and anisotropic permeability. DA was applied
Recovery factor
Scaling groups
using Buckingham Pi theorem resulting in twelve dimensionless groups. In IA, using the governing equations and
non-dimensionalization procedure, it turned out that twelve scaling groups are required to reach the similarity of
recovery curves in two different scales (i.e. core and field scales). The resultant scaling groups based on both
methods were then compared. To further verify these scaling groups, recovery curves of three scenarios with
different parameters and the same scaling groups were also compared. Several sensitivity analyses were also
performed to discern the impact of each newly developed scaling group to change the breakthrough recovery
factor (BRF). Through the sensitivity analysis, one scaling group was proved to be redundant in changing BRF,
whereas one was repetitive. Therefore, it was found that ten scaling groups were sufficient for scaling LSWF from
core to field scale. The results showed that deviation of the system from Vertical Equilibrium (VE) and scaling
groups related to mobility ratio, capillary forces and gravity segregation dominate the total sweep efficiency (i.e.
vertical sweep efficiency) of LSWF.

1. Introduction mechanisms contributing to fractional oil recovery (Austad et al., 2008;


Bagci et al., 2001; Gupta and Mohanty, 2011; McGuire et al., 2005; Patil
Energy demand, which is largely provided by oil/gas resources, rises et al., 2008; Seccombe et al., 2010; Webb et al., 2005; Yousef et al.,
particularly in the recent years (www.enerdata.com, 2018) whereas the 2011; Zhang et al., 2007). However, there is no consensus over the
proved reservoirs are not infinite in a manner that the investigated hy­ dominancy of a specified mechanism over the others. Moreover,
drocarbon resources increased only 60% over the last two decades modeling of LSWF has not yet been developed sufficiently as well as the
(British Petroleum, 2016). On the other hand, insufficient oil production dominant mechanisms which are still under scrutiny.
during primary and secondary recovery causes a large part of the initial Despite the uncertainties about the dominant mechanisms in LSWF,
oil in place to be trapped in the reservoir (Lake, 1989). Therefore, the some models have been developed and verified with experimental and
implementation of EOR methods to recover the remaining oil would field results (Al-Shalabi and Sepehrnoori, 2017). Jerauld and Lin (2008)
look increasingly indispensable (Al-Shalabi and Sepehrnoori, 2017). A proposed a method to simulate the behavior of the reservoir system
number of EOR techniques have been used to cope with the difficulties during LSWF. In their study, the correlation between the relative
of moving the rest of oil, trapped in the reservoir. These approaches are permeability and capillary pressures was proportional to the deviation
thoroughly categorized and discussed by Alvarado and Manrique of salinities from thresholds (Jerauld and Lin, 2008). Zeinijahromi et al.
(2010). (2013) developed a mathematical model to simulate formation damage
LSWF (low salinity waterflooding) is one of the tangible methods at related to LSWF using polymer modules, in which the residual resistance
both laboratory and field scales for the purpose of enhanced oil recovery factor was used as a criterion to obtain the reduction of water relative
(Al-Shalabi and Sepehrnoori, 2017). In sandstones and carbonates, permeability, i.e. formation damage. Dang et al. (2016) proposed a
experimental investigations have established the efficiency of different mechanistic LSWF model which considered the effects of ion exchange

* Corresponding author.
E-mail address: mriazi@shirazu.ac.ir (M. Riazi).

https://doi.org/10.1016/j.petrol.2020.106956
Received 16 March 2019; Received in revised form 9 January 2020; Accepted 14 January 2020
Available online 21 January 2020
0920-4105/© 2020 Elsevier B.V. All rights reserved.
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

and geochemical reactions in oil mobilization due to the enhanced water and IA persuaded researchers to use them simultaneously. In fact, using
spreading at low salinity. While validating their model, it was observed both methods would help to identify all scaling groups which physically
that in the coreflood simulation, the use of Ca2þ equivalent fraction on can predict the behavior of the system (Zhou and Yang, 2017).
the exchanger as the interpolant provides a good match with the Flow in the porous media would usually be controlled by different
experimental data. Furthermore, Korrani and Jerauld (2018) used a mechanisms while, in the event of implementing EOR methods, the
surface-complexation-based model to predict wettability alteration in degree of complexity is expected to be much higher. Therefore, scaling
both sandstones and carbonates. They modelled the wettability alter­ under such circumstances is not as simple. To scale these systems, usu­
ation with zeta potentials using stability number, an interpolant equals ally two solutions are proposed whether the governing equations are
to the ratio of electrostatic to van der Waals forces (Korrani and Jerauld, available (Andersen et al., 2017) or not (Berawala et al., 2019; Nygård
2018). The fractional flow of low salinity polymer flooding, including and Andersen, 2019).
wettability alteration, was investigated by Khorsandi et al. (2017). They Buckingham (Pi) theorem is a formalization of the Rayleigh’s
considered water viscosity as a strong function of salt and polymer method in DA. This theorem states that the number of (Pi) terms after
concentrations, whereas, relative permeabilities were controlled by performing DA is equal to the difference between the number of perti­
water saturation and salt concentration. In addition, the interpolation nent parameters and the maximum number of fundamental dimensions.
parameter was attained using cation exchange (Khorsandi et al., 2017). Moreover, Buckingham theorem is a way of defining dimensionless
Further studies performed by Al-Ibadi et al. (2018a) using a fractional groups, however, the generated Pi terms may not be unique (Zohuri,
flow model including the salt diffusion. In their study, they used effec­ 2015). Hence, the selection of correct repetitive parameters is impera­
tive salinity range to obtain weighting function to interpolate between tive. In many scaling processes, only [M], [L] and [T] are proposed as
high-salinity and low salinity states. Al-Ibadi et al., 2018 also used a 1-D fundamental dimensions. In a novel method, proposed by Novakovic
model with no gravity effect. They finally proposed a method to obtain (2002) , [-] is also added to fundamental dimensions. It would help to
the recovery curves without any need for simulation. find out scaling groups specially those which are inherently dimen­
Despite the usefulness of LSWF, under certain circumstances, this sionless. This indicates that it would be more helpful if the scaling
method similar to the other EOR methods has its own shortcomings such groups are compared to those obtained from IA.
as low recovery factor and high water production in a way that could In LSWF, the final form of governing equations is approximately
make it impractical. Thus, to examine the deficiencies of different EOR similar to waterflooding, which can help in scaling of LSWF. As stated
approaches in a typical reservoir, it is convenient to inspect them at before, there are numerous studies related to scaling of waterflooding.
small scale, then, scaling their results up from lab to field scale. One such comprehensive study was performed by Shook et al. (1992) for
Primarily, scaling was used in reservoir engineering in 1942 to scale IA of waterflooding. In their research, a 2-D vertical reservoir model
fluid flow around a well (Leverett et al., 1942). Then, in 1951, dimen­ with homogeneous and isotropic permeability was applied to obtain five
sionless groups were used to investigate the effect of oil displacement scaling groups. Provided this short literature, it is evident that the
with water in a granular sand pack (Engelberts and Klinkenberg, 1951). scaling groups have not thoroughly been developed for LSWF. Accord­
In 1955, Croes and Schwarz using Buckley Leverett and Dietz’s theory ingly, in this study, dimensionless terms for LSWF would be derived
represented five scaling groups for waterflooding. They used the results based on both DA and IA. Furthermore, the gravity forces as well as
of their own experiments to validate them (Schwarz and Croes, 1955). viscous and capillary forces have been considered. The effect of all
Since then, dimensionless groups were widely used to scale laboratory scaling groups on LSWF have also been investigated using breakthrough
results to field scale (Craig et al., 1957; Geertsma et al., 1956; Gharbi recovery factor. To do so, the proposed model by Jerauld in 2008 was
et al., 1996, 1998; Gharbi, 2002; Jadhawar and Sarma, 2008; Perkins applied for LSWF simulations in the present study. Finally, the sensi­
and Collins, 1960; Shook et al., 1992; Sorbie et al., 1994; Thomas et al., tivity of BRF with respect to each scaling group would be examined.
2000; Carpenter et al., 1962; Greenkorn, 1964; Greenkorn et al., 1965;
Davis and Jones, 1968; Pujol and Boberg, 1972; Parsons, 1977; Hirasaki, 2. Theory and simulation
1980). For instance, van Daalen and van Domselaar (1972) focused on
how to scale fluid flow models with different geometries using inspec­ 2.1. Numerical simulation
tional analysis. They proposed effective aspect ratio instead of the ratio
length/height. They also introduced dip number and a new form gravity In this study, a commercial simulator (E100, GeoQuest, 2010), is
number (van Daalen and van Domselaar, 1972). Furthermore, Shook used to apply numerical simulation in a Cartesian system and
et al. (1992) proposed a methodology to attain independent scaling block-center coordinate. In what follows, various aspects of the nu­
groups for waterflooding. merical simulation such as definition of geometry, assumptions and
In spite of many investigations for scaling of waterflooding, none­ simplifications, governing equations, and finally the scaling groups are
theless only a few studies are available with respect to LSWF scaling discussed.
groups. For instance, Al-Ibadi et al. (2018) used the effective salinity
range as a scaling group. On the other hand, there are some scaling 2.2. Geometry
groups related to waterflooding which could also be used for LSWF.
However, these should be examined in terms of their applicability. To evaluate vertical sweep efficiency and also for grid generation, a
Zapata and Lake (1981) proposed a scaling procedure to obtain two-dimensional porous medium with homogeneous and anisotropic
dimensionless groups in a 2-D porous media without capillary and permeability is considered (Fig. 1). This 2-D porous medium intersects
gravity effects. In fact, they focused on viscous vertical equilibrium in the horizontal plane with a tilted angle equal to α. The low-salinity water
two-phase flow. They also developed two mobility terms, i.e. mobility is injected through the porous media in the x-axis direction, while the z-
ratio across the oil bank and mobility ratio across the chemical front. axis is at right angle to the principle direction of the flow. In a two-
Furthermore, they used permeability contrast for permeability ratio of dimensional medium, vertical displacement efficiency controls the
layers. total sweep efficiency. Vertical sweep efficiency is, in turn, influenced by
Generally speaking, scaling can be categorized into two approaches: mobility ratio, capillary forces, gravity segregation due to the density
Dimensional Analysis (DA) and Inspectional Analysis (IA) (Geertsma difference and horizontal or vertical permeability changes (Bahadori,
et al., 1956). Non-uniqueness of dimensionless groups aside, IA is 2018). In the present reservoir model, all these factors except hetero­
required to ensure that the scaling groups are physically meaningful geneity in permeability are considered.
(Craig et al., 1957). However, the number of scaling groups obtained
using IA is not reliable and should be validated with DA. Inabilities of DA

2
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Fig. 1. A typical two-dimensional sketch of LSWF.

2.3. Assumptions for LSWF simulation


krw ¼ F1 � kLS
rw ðSwD Þ þ ð1 F1 Þ � kHS
rw ðSwD Þ (1-1)
As explained in the preceding sections, different methods are sug­
gested to model LSWF. In the present study, the proposed method by kro ¼ F1 � kLS
ro ðSwD Þ þ ð1 F1 Þ � kHS
ro ðSwD Þ (1–2)
Jerauld and Lin (2008) is followed to simulate LSWF. For simplification,
the following assumptions are made: Pcow ¼ F1 � PLS
cow ðSwD Þ þ ð1 F1 Þ � PHS
cow ðSwD Þ (1–3)

F1 ¼ f ðCs Þ (1–4)
� 2-D single porosity (homogeneous) and single permeability media
(homogeneous and anisotropic).
SwD ¼ ðSw Swco Þ=ð1 Swmax Swco Þ (1–5)
� Two-phase flow of water and oil in the porous media (no gas exists in
the flow).
� In different scenarios, the reservoir model’s system is assumed hor­ � In addition, initially, in black oil simulation, the following relation­
izontal or tilted. ships are used to attain saturation end-points (GeoQuest, 2010).
� In simulations, salt is an individual constituent (i.e. the composition � HS
Swco ¼ F LS LS
1 Swco þ 1 F LS
1 Swco (1–6)
of salt is not considered), which is used to estimate other parameters
such as saturation and pressure at each time-step. � HS
Swmax ¼ F HS LS
1 Swmax þ 1 FHS Swmax (1–7)
� Experimental results showed that with small salinity changes, the 1

aqueous phase parameters do not change considerably (Kwak et al.,


2005; Osif, 1988). Therefore, in the present study, it is assumed that � In all scaling processes, it is assumed that the initial and residual
saline water properties (i.e. density and viscosity) remained intact saturations in high-salinity and low-salinity relative permeabilities
with salinity changes. are set to specified values and changes in saturation end-point would
� Salinity is modelled using an effective salinity range and saturation not be considered. In fact, high-salinity and low-salinity relative
functions’ interpolation parameter (F1) to gradually switch from permeability curves in all scenarios would have the same corre­
high to low salinity relative permeabilities. In all previous models for sponding saturation end-points. Therefore, scaling groups related to
LSWF, attaining this parameter is an important point of consider­ saturation would be neglected.
ation, in which, in most cases its functionality should be examined.
Initial and injection salinities are selected within salinity range, high 2.4. Governing equations
salinity (CHS LS
s ) and low salinity (Cs ) concentrations. Thus, salinities
outside of the extreme boundaries cannot be used. The first step to form the scaling groups, essentially in IA, is to define
� Interpolation parameter (i.e. F1 in Equations (1–1)) ranges from zero the governing equations (Shook et al., 1992). Governing equations in
to one. It only follows the salinity of a cell. Hence, only its trend LSWF are explained below.
versus salinity is important. Mass conservation of salt component in water phase:
� Parameters of capillary pressure and relative permeability curves are
strong function of salinity which are defined as follow (Jerauld and ∂ðCs Sw Þ ∂ðCs uwx Þ ∂ðCs uwz Þ
φ þ þ ¼0 (2)
Lin, 2008): ∂t ∂x ∂z
Mass conservation of water phase:

3
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

∂Sw ∂uwx ∂uwz simulations. Skjaeveland et al. (1998) used the following equations for
φ þ þ ¼0 (3)
∂t ∂x ∂z saturation functions:
Mass conservation of total fluids in the system: kro;ow ¼ kLðHÞS ð1 SwD Þ
3þ2aw
(11–1)
ro
∂ðuwx þ uox Þ ∂ðuwz þ uoz Þ
þ ¼0 (4) krw;ow ¼ kLðHÞS S1þ2a w 2
SwD (11–2)
∂x ∂z rw wD

Darcy equation: �
(11–3)
2
� � kro;ww ¼ kLðHÞS
ro 1 S1þ2a
wD
w
ð1 SwD Þ
krw ∂Pw
uwx ¼ kx þ ρw g sin α (5–1)
μw ∂ x krw;ww ¼ kLðHÞS
rw S3þ2a
wD
w
(11–4)
� �
kro ∂Po cwd krw;ww cod krw;ow
uox ¼ kx þ ρo g sin α (5–2) kLðHÞS ¼ (11–5)
μo ∂ x rw
cwd cod
� �
kro ∂Po cwd kro;ww cod kro;ow
uoz ¼ kz þ ρo g cos α (5–3) kLðHÞS
ro ¼ (11–6)
μo ∂z cwd cod
� �
krw ∂Pw CLðHÞS CLðHÞS
uwz ¼ kz þ ρw g cos α (5–4) Pc ¼ � w
�aw þ � o �ao (11–7)
μw ∂z Sw Swco So Sor
1 Swco 1 Sor
By assuming oil as a non-wetting phase, capillary pressure would
then be defined as follows: For simplicity, in the following scaling processes, the area below the
saturation functions versus saturations were used as representative of
Pc ¼ P o Pw (6)
the saturation functions curves. Therefore, only by changing the vari­
At the inlet and outlet of the system, the following equations can be ables in Table 1, these areas were changed.
applied.
Z
1 H 2.5. Scaling groups
uwx dz ¼ uT at x ¼ 0; 8t (7)
H 0
As already mentioned, scaling groups can be derived from both IA
Pw ¼ Po ¼ Pwf þ ðρg cos αÞðH zÞ at x ¼ L; 8z; t (8) and DA. In DA, scaling groups may have no physical meaning. This is the
case when there is no certainty about the applicability of scaling groups;
To derive the average density, the following equation is used. however, the number of scaling groups is reliable. In other words,
Z H dimensional analysis cannot guarantee that the obtained scaling groups
uwx dz are the best selected. This is because using different repetitive parame­
ρ ¼ ρo þ Δ ρ 0 at x ¼ L; 8t (9) ters could make different scaling groups. Under such circumstances, IA
HuT
provides much stronger judgement about the meaningfulness of scaling
where Δρ ¼ ρw ρo and uT ¼ uwx þ uox . groups as IA is primarily based on the governing equations.
Water and oil velocities at the vertical boundaries are as the
following: 2.5.1. Dimensional analysis (Buckingham’s (Pi) theorem)
At first, the Buckingham’s (Pi) theorem scaling groups need to be
uwz ¼ 0 at z ¼ 0; 8x; t (10–1)
defined. In this case, initially the pertinent parameters have to be
introduced which were as follows.
uwz ¼ 0 at z ¼ H; 8x; t (10–2)
HS HS LS LS
f L; H; φ; kz ; kx λw ; kx λo ; kx λw ; kx λo ; Cs ;
uoz ¼ 0 at z ¼ 0; 8x; t (10–3) LS HS � (12–1)
F1 ; Δρow g cos α; ρw g cos α; Pc ; Pc ; uT ; t ¼ 0
uoz ¼ 0 at z ¼ H; 8x; t (10–4)
Four fundamental dimensions were chosen to attain the scaling
groups ([M], [L], [T] and [-]). The choice of the repeating variables P1,
2.4.1. Saturation functions
P2, …, Pk should be such that they include all the K dimensions used in
Relative permeability and capillary pressure functions are generated
the problem (Zohuri, 2015). Based on the preliminary inspection, the
using the method proposed by Skjaeveland et al. (1998). In numerical
repetitive parameters are as follows:
simulation, it is assumed that the system alters from oil-wet to water-wet
(Al-Shalabi and Sepehrnoori, 2017). Moreover, as there is not any
Table 1
consensus on the selection of relative permeability curves from oil-wet
Constants and variables used in equations (11-1) – (11-7).
to water-wet conditions, then all possible changes in relative perme­
ability should be considered. Furthermore, in the event of formation No. Parameter Value No. Parameter Value
(unit) (unit)
damage, there is no certainty about the final relative permeability
curves. In 2014, Al-Shalabi et al. showed that in history matching pro­ 1 aw ð Þ 0.20 6 kLðHÞS
rw;ww ð Þ Varies in
different
cess of LSWF at core scale, tuning both relative permeability end-points
scenarios
and Corey’s exponents is indispensable (Al-Shalabi et al., 2014). 2 Varies in 7 0.6
CLðHÞS
w ðpsiÞ kLðHÞS
ro;ow ð Þ
Therefore, it can be stated that to generate water-wet and oil-wet rela­ different
tive permeability curves, both of them should be changed. Thus, in scenarios
LSWF simulations, it was attempted to adhere to these principles, 3 ao ð Þ 0.10 8 kLðHÞS 0.4
rw;ow ð Þ
although a change in saturation end-points from high-salinity to 4 CLðHÞS
o ðpsiÞ 0.001 9 cwd ð Þ 1.000
low-salinity condition is assumed. Table 1 provides different constants 5 kLðHÞS
ro;ww ð Þ Varies in 10 cod ð Þ 0.001
and variables used for constructing relative permeability and capillary different
scenarios
pressure curves. These constants would be used in the numerical

4
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

HS �
L; φ; kx λw ; uT HS HS
kx kro PCow PCow
LS

NB ¼ (23)
Then, each scaling group can be obtained by combination of these uT Lμo
parameters and one from the remaining parameters (here Ω such as H;kz ; �
HS LS
HS LS LS LS HS
kx λo ;kx λw ;kx λo ;Cs ;F1 ;Δρow g cos α; ρw g cos α;Pc ;Pc ;t). For example, for NPc ¼
PCow PCow
(24)
a specific Pi term, one would have:
HS
PCow
Y HS �c
¼ La φb kx λw udT Ωe It should be pointed out that some scaling groups are called as “G” in
x
Appendix A which are proposed in this study but initially no firm
Which Ω is one of the remaining parameters, whereas a, b, c, d and e judgement could be made whether they are independent or dependent.
are the exponents that should be obtained in order to provide dimen­ Once their independency is established then a uniform name would
sionless scaling groups. instead be used.
Finally, by executing the Buckingham’s (Pi) theorem, all potential In the above dimensionless groups, Equations (13) and (14) show the
scaling groups were obtained as follows: saturation functions’ interpolation parameters groups, Equations (15)–
� . . .pffiffiffiffi HS . LS HS . LS HS . HS (17) are related to mobility of fluids in porous media, Equation (18)
f L H; uT t Lφ; L kz ; λw λw ; λw λo ; λw λo stands for VE, Equations (19)–(21) show the gravity forces, Equation
HS � HS �
; Cs ; F1 ; kx λw Δρow g cos α uT ; kx λw ρw g cos α uT (12–2) (22) represents the dip of reservoir and Equations (23) and (24) are
HS HS � HS LS � � related to capillary forces. Meanwhile, Nρ and Nρ2 seems to be dimen­
; kx λw Pc uT L; kx λw Pc uT L ¼ 0 sionally dependent, whereas using Nρ Nρ2 ¼ 1 explicitly confirms that
these two scaling groups are dependent on each other. Thus, in the
2.5.2. Inspectional analysis (IA) following sections, only one of them was applied. After reviewing the
In this section, initially governing equations in LSWF are presented. scaling groups using both IA and DA, in the following section, they
Then, the proposed procedure to derive scaling groups is stated before would be compared to discern the inaccuracy of scaling groups obtained
selecting the final scaling groups. from DA in relation to IA for identification of scaling groups.

2.5.3. Derivation of scaling groups 2.5.4. Comparison of the scaling groups


As shown in the preceding section, 18 parameters (L;H;∅ ;kz ;kx ;Cs ;F1 ; Generally speaking, as stated before, the scaling groups which would
uT ; ρo ; ρw ; α; λHS LS HS LS HS LS
w ; λw ; λo ; λo ; Pc ; Pc ; Pwf ; t) are involved in the process of be obtained using IA are more reliable than DA since they are based on
LSWF for the defined system. To simplify the system, it was assumed that governing equations. As can be seen from Table 2, some scaling groups
relative permeability curvature and residual saturation remained intact. attained using DA were similar to those in IA. Worthwhile also to say
Using non-dimensionalization, some dependent scaling groups were also that some other groups in IA can be attained by applying different
derived. The procedure of attaining the scaling groups using inspec­ mathematical operations to Buckingham Pi terms.
tional analysis have been shown in Appendix A. After attaining the
number of scaling groups, it was attempted to find the suitable scaling 3. Results and discussion
groups. Finally, the following scaling groups were proposed for scaling
of LSWF. In this section, firstly the scaling groups would be validated using
parameters in their most diverse range of variations but of course for the
Gint ¼ F1 (13)
same scaling groups. Then, the sensitivity analysis of each scaling group
at inter-well scale would be reviewed.
Cs CLS
CSD ¼ s
(14) It should be pointed out that numerical dispersion is important
CHS CLS
s s which takes into account the deviation of simulation results from the
LS reality. There are several reasons for this. For instance, as grid size and
LS
M oHS ¼
kro
(15) time-steps would not have appropriate resolution, these may then affect
the simulation results. Therefore, it is important to have proper resolu­
HS
kro
tion to achieve acceptable forecasts. In addition, Sorbie and Mackay
krw
LS
(2000) investigated the effects of mixing of injected, connate and aquifer
(16)
LS
M wHS ¼ HS
krw

HS
Table 2
krw μo Comparison of the resultant scaling groups using Buckingham Pi theorem and
M HS ¼ (17)
HS
kro μw IA.
No. Buckingham Pi theorem IA
sffiffiffiffi
L kz 1 L=H L tan α=H
RL ¼ (18)
H kx 2 uT t=Lφ uT t=Lφ
pffiffiffiffi pffiffiffiffi pffiffiffiffiffi
3 L= kz ðL kz Þ=ðH kx Þ
ρo 4 HS LS LS HS
Nρ ¼ (19) λw =λw kro =kro
Δρ
5 HS LS
λw =λo
LS HS
krw =krw
ρw 6 HS HS HS HS
Nρ2 ¼ (20) λw =λo ðkrw μo Þ=ðkro μw Þ
Δρ 7 Cs CsD

HS
8 F1 F1
kx kro Δ ρg cos α 9 HS HS
Ng ¼ (21) kx λw Δρow g cos α=uT kx λw Δρow g cos α=uT
uT μo
10 HS
kx λw ρw g cos α=uT ρw =Δρ

L 11 HS HS
kx λw Pc =uT L
HS
kx λo ðPCow
HS LS
PCow Þ=ðuT LÞ
Nα ¼ tan α (22)
H 12 HS LS
kx λw Pc =uT L
HS
ðPCow
LS HS
PCow Þ=PCow

5
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

brines in field scale. To prevent numerical dispersion, they used Peclet Table 3
number to obtain the appropriate grid sizes for finite difference simu­ Input data of different scenarios used for validation of scaling groups.
lations. Numerical dispersivity can be attained using the following No. Parameter Scenario No.
expression.
1 2 3
� �
δx Vδt 1 LðftÞ 300 100 10
αL ¼ (25–1)
2 2 2 HðftÞ 30 10 1
3 kz ðmDÞ 10 5 1
where the Peclet number for single-phase flow in porous media is 4 kx ðmDÞ 200 100 20
defined as follows. 5 HS 0.5 0.6 0.7
krw
L 6 LS 0.5 0.6 0.7
NPe ¼ (25–2) krw
αL 7 HS 0.4 0.5 0.6
kro
Sorbie and Mackay (2000) also stated that for a sufficiently small 8 LS 0.50 0.63 0.75
kro
time step, the level of numerical dispersion is entirely due to the spatial
9 FLS 0 0 0
term. 1
10 FHS
1
1 1 1
δx 11 CLS 0.3 0.2 0.1
αL ¼ (25–3) s
2 12 CHS 30 20 10
s

In the present study, it was tried to follow reverse proportionality 13 Cinj


s
0.3 0.2 0.1
between the injection velocity and time steps. However, due to small 14 Cinit
s
30 20 10
time-steps, the time term in Peclet number can be neglected. 15 μo ðcPÞ 4.0 2.7 1.7
16 μw ðcPÞ 0.60 0.40 0.25

3.1. Validation of scaling groups 17 α 0 0 0


� �
18 3.1E- 2.8E- 1.0E-
uT ft=
s 08 8 9
As shown in Table 3 and Table 4, to ensure the appropriateness of the
19 φ 0.05 0.10 0.20
identified scaling groups, different pertinent parameters (row 1–23,
20 PLS
Cow ðpsiÞ
0 0 0
Table 3) were used for three scenarios as characterized in Table 3 hence
21 PHS 90 30 3
the same scaling groups could be obtained (see Table 4). In addition to Cow ðpsiÞ
22 ρo ðlbm =ft3 Þ 56.0 56.0 56.0
pertinent parameters (row 1–23), Table 3 also presents the values of
23 ρw ðlbm =ft3 Þ 67.0 67.0 67.0
non-scaled parameters (row 24–30). Then, the uniqueness of recovery
curve versus dimensionless time was checked. In Fig. 2, recovery was 24 Initial pressure (psi) 3000 3000 3000
25 Bottom-hole pressure of production well (psi) 2950 2950 2950
plotted versus log(tD) for the three scenarios. As can be seen, a proper 26 Time step (day) 0.1 0.1 0.1
match of three recovery curves confirms how the proposed scaling 27 SLS 0.3 0.3 0.3
wco ( )
groups are suitable for LSWF. 28 SHS 0.3 0.3 0.3
wco ( )
29 SLS
or ( )
0.2 0.2 0.2
3.2. sensitivity analysis 30 SHS
or ( )
0.3 0.3 0.3

The sensitivity analysis yields important results about the behavior


of a representative 2-D system in the LSWF process. Initially, the ability
Table 4
of different scaling groups in changing the behavior of the system has to
Scaling groups for different scenarios used for validation of scaling groups.
be evaluated. To do so, the sensitivity of each scaling group was char­
No. Scaling Group Scenario No.
acterized in terms of its impact on the breakthrough recovery factor as it
changed while keeping the other scaling groups intact. The details of this 1 2 3
procedure is provided in Appendix B (Tables 9 and 10). 1 Gint ¼ F1 0.5 0.5 0.5
As it was also mentioned earlier, for the present reservoir model as 2 LS LS 1.25 1.25 1.25
described in section 2, vertical sweep efficiency was the sole controlling kro
MoHS ¼ HS
factor of total sweep efficiency. Consequently, by presuming a homo­ LS
kro
3 1.0 1.0 1.0
geneous permeability, the total sweep efficiency would be dominated by
LS
krw
MwHS ¼ HS
mobility ratio, capillary forces and gravity segregation (Bahadori, krw
4 HS 6.8 6.7 6.7
2018). The recovery factor at breakthrough time (BRF) was considered MHS ¼ rw
k μo
HS μ
as the criterion for examination of sensitivity analysis of the resultant ksro ffiffiffiffiffi
w
5 L kz 2.23 2.23 2.23
scaling groups. In this regards, a tracer was used in the injection well. As RL ¼
H kx
its effect was observed in the production well, recovery factor was 6 ρ 6.1 6.1 6.1
Nρ ¼ w
recorded. At reservoir condition, capillary forces are not as critical as Δρ
7 L 0 0 0
gravity or viscous forces. Therefore, initially, for simplification of Nα ¼ tan α
H
sensitivity analysis, zero capillary pressure condition (for both high 8 HS
kx kro Δρg cos α 4.5 4.4 4.4
Ng ¼
salinity and low salinity water) was assumed. Then, for sensitivity uT μo
9 HS HS LS 14 14 14
analysis of both dimensionless numbers related to the capillary forces, NC ¼
kx kro ðPCow PCow Þ
uT Lμo
the effect of non-zero capillary numbers were also considered. It should 10 HS LS 1 1 1
ðP P Þ
be noted that in sensitivity analysis process, the gridding and time-steps NPc ¼ Cow HS Cow
PCow
in all scenarios were constant. 11 Cinj CLS 0.0 0.0 0.0
Cinj s s
As for different scaling groups, the trend of their effects on break­ SD ¼ HS
Cs CLS
s
12 Csinit
CLS 1.0 1.0 1.0
through recovery factor was observed in a 2-D system. These groups can Cinit
SD ¼ HS
s
Cs CLS
be extended to 3-D systems by considering all pertinent forces (i.e. s

viscous, capillary and gravitational forces) with respect to LSWF in

6
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Fig. 2. Validation of scaling groups using recovery factor vs. Log(tD) for three different scenirios as defined in Table 3.

Fig. 3. Relative permeability and interpolant data: a) High-salinity and low-salinity relative permeabilities b) Interpolant parameter (F1) vs dimensionless
salinity (CSD).

7
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

qffiffiffi
Table 5
3.2.1. Effect of effective aspect ratio (RL ¼ L
kx ) on breakthrough
kz
Numerical simulation data for sensitivity analysis of the proposed scaling groups H

at 2-D reservoir condition. recovery factor


No. Parameter Value
Vertical equilibrium (VE) is a state of flow where the highest cross­
flow is met and the sum of fluid-flow driving forces at right angle to the
1 Number of grid blocks (Nx, Ny, Nz) 500, 1, 30
bulk flow direction is zero. VE is usually met where high RL exists (Lake
2 Grid block size (Δx, Δy, Δz) (ft) 1, 10, 1
3 Length (ft) 500 et al., 2014). In other words, this dimensionless number determines the
4 Height (ft) 30 tendency of the system to reach the VE (Shook et al., 1992). As it can be
5 Vertical permeability (mD) 50 seen from Fig. 4, with increasing RL more than 5.27, the changes in BRF
6 Horizontal permeability (mD) 250 with gravity number does not change noticeably. In fact, it shows that
7 μo (cP) 3
the forces perpendicular to the flow path cannot alter the behavior of the
8 μw (cP) 0.5
system. However, at lower gravity number values, this separation be­
9 ρo (lbm/ft3) 56.0
tween the curves is not clearly distinguishable. This implies that BRF in
10 ρw (lbm/ft3) 67.0
LSWF is in control of flow regime as well as VE.
11 Porosity (fraction) 0.05
For sensitivity analysis of the following scaling groups, RL was
12 Csinj (lbm/STB) 0.35
13 Csinit (lbm/STB) 20 selected in a region where vertical equilibrium would exist (RL ¼ 7.45).
14 Time step (day) 0.1 It helped in two directions, one in reaching the maximum capability of
15 Top of reservoir (ft) 8000 crossflow and the other in getting closer to the reservoir conditions.
16 Initial pressure (psi) 3000
17 Bottom-hole pressure of production well (psi) 2950
3.2.2. Effect of dip number (Nα ¼ HL tanα) on breakthrough recovery factor
In up-dip injection, the gravity force can improve displacement ef­
porous media; nonetheless, their impact should again be scrutinized. ficiency where the displacing fluid has lower density compared to the
There are several cases which performed for 2D and 3D scaling analyses displaced fluid (Bahadori, 2018). The dip number is an appropriate tool
of waterflooding and showed similar results (Novakovic, 2002; Shook to examine the magnitude of this effect. Fig. 5 shows the effect of Nα on
et al., 1992). BRF for different flow regimes. As it can be seen, by changing gravity
For sensitivity analysis, the relative permeability and interpolation number as result of dip, the low salinity water can sweep oil better than
parameter curves are shown in Fig. 3. In addition, the other numerical non-dipping surfaces, as gravity is dominant. In other words, with
data points are tabulated in Table 5. For sensitivity analysis of different increased dip angle, the gravity effects diminished in a way that would,
scaling groups though a change in scaling groups should be examined. in turn, improves the sweep efficiency. This would imply that in tilted
In the following sections, initially it would be assumed that capillary reservoirs, flooding with lower viscous forces would help in producing
does not exists in the present reservoir model, but finally, its effects were more oil before breakthrough with postponing fingering. Accordingly, in
examined. This was followed in the sensitivity analyses for two reasons. LSWF, it can be inferred that gravity segregation in up-dipping reser­
Firstly, the gravity force is important at reservoir conditions. It would voirs would improve vertical displacement efficiency.
then be a good choice for discerning dominant scaling groups. Secondly,
as viscous, capillary and gravity forces exist at reservoir conditions, then 3.2.3. Effect of density ratio (Nρ ¼ Δρ)
ρw
on breakthrough recovery factor
to investigate their impacts on the flow behavior, it is important to keep The effect of density ratio in both dipping and non-dipping reservoirs
one of them zero before comparing the competition between the other showed that it has no profound impact on BRF (Fig. 6) which in
two. To achieve this aim, initially in evaluation of different scaling compliance with what Shook et al. (1992) have also reported for
groups, gravity and viscous effects are considered and finally, after waterflooding. The ratio of the densities has also no noticeable influence
reviewing the effects of these forces, at different specified Ng capillary on the vertical sweep efficiency, whereas, the difference between
effect is also examined.

Fig. 4. Breakthrough recovery factor versus Ng for different RL.

8
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Fig. 5. Breakthrough recovery factor versus Ng for various Nα .

densities which shows its effect in gravity number is an important factor. 3.2.7. Effect of effective salinity range (CSD ¼
Cs CLS
s
) on breakthrough
Therefore, it seems to be right to ignore this dimensionless number from CHS
s CLS
s

recovery factor
the pertinent scaling groups.
This number has previously been used by Attar and Muggeridge
LS LS (2016) and Al-Ibadi et al. (2018). As shown in the preceding sections,
3.2.4. Effect of (MoHS ¼ ) on breakthrough recovery factor
kro
HS
kro
the dimensionless number related to the salinity includes a range of
Increasing this number leads to decreased mobility ratio in the pre­ salinity from initial condition, as only high salinity water exists to the
sent reservoir model. Therefore, it is anticipated to postpone break­ final condition when low salinity water swept the porous media. In this
through time as result of pushing the flooding process towards respect, two extremes were used as representatives of salinity in the
appropriate mobility ratios. As can be seen from Fig. 7, at low gravity system.
LS
numbers, increasing MHS
o results in a sharp increase in BRF. On the CInj
3.2.7.1. Effect of injection salinity number (CInj ) on
s CLS
SD ¼
s
LS
contrary, at higher gravity numbers, increasing Mo has partial effect on
HS CHS
s CLS
s

breakthrough recovery factor. It implies that at higher gravity numbers, breakthrough oil recovery factor
the ratio of low salinity to high salinity oil relative permeability has Injected water dimensionless salinity controls the residual oil satu­
minor effect on BRF, whereas at lower gravity numbers, rise in this ration. On the other hand, as result of its effect on F1, it would control
number results in higher breakthrough recovery factor. This indicates the mobility ratio. Thus, as this number increases then the tendency of
the effect of favorable mobility ratio with increasing oil mobility during the system to finger flow would be lower. Therefore, as breakthrough
LSWF, would result in higher breakthrough recovery factor. time postponed, higher BRFs are expected (Fig. 10).

3.2.7.2. Effect of initial salinity number (CInit ) on breakthrough


CInit CLS
SD ¼
LS LS s s
3.2.5. Effect of (MwHS ¼ ) on breakthrough recovery factor
krw CHS CLS
HS s s

LS
krw
oil recovery factor
Similar to MoHS , the effect of this scaling number is higher at low CInit
SD increases the tendency of the system towards high-salinity water
gravity numbers. In fact, at high gravity numbers, this scaling number (saturation functions and end-point saturations). Thus, the higher this
has lower effects on BRF (Fig. 8). On the contrary, at low gravity number, the lower breakthrough recovery factor would be expected
LS
numbers, by increasing MwHS , breakthrough recovery factor is expected to (Fig. 11).
decrease. This implies that higher values of this number stimulates the
system towards unfavorable mobility ratios, resulting in lower BRF. 3.2.8. Effect of F1 on breakthrough recovery factor
As it was mentioned in the previous sections, F1 as a function of
HS salinity could only vary between zero and one. It also serves as inter­
3.2.6. Effect of high salinity water mobility ratio (MHS ¼ ) on
krw μo
HS
kro μw polant between high-salinity and low-salinity saturation functions.
breakthrough recovery factor Therefore, as salinity number, it can affect both end-point saturation and
The mobility ratio of high salinity water (i.e. initial mobility ratio) relative permeability. In this section, the effect of area below the curve
profoundly controls the breakthrough recovery factor (Fig. 9). It is clear of F1 versus CsD (average F1) is investigated. In fact, this parameter is
that in mobility ratios greater than one, the breakthrough happens always defined as the linear function of salinity, in which, this may
earlier than other situations when the mobility ratio is less than one. As sometimes be an error especially at real circumstances. Therefore, this
it can be seen from Fig. 9, mobility ratio of high salinity water is a section discusses how nonlinearity of this parameter can alter the
dominant factor; which controls the breakthrough recovery factor even results.
when the gravity dominated flow exists. It is obvious that F1 would have a key role in defining end-point
saturations and relative permeabilities in each time-step. It can affect

9
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Fig. 6. Breakthrough recovery factor versus density ratio (Nρ ) for different Ng in a) non-dipping and b) dipping surfaces.

the results in two distinct ways. Firstly, it can help the definition of 3.2.9. Effect of modified capillary number (NC ¼
kx kHS HS LS
ro ðPCow PCow Þ
) on
mobility ratio with known salinity in each time-step. Secondly, F1 would
uT Lμo
breakthrough recovery factor
affect the end-point saturations in a way which would result in different
This dimensionless number shows the changes of capillary pressure
ultimate recovery factors based on extremes of dimensionless salinities.
between low salinity and high salinity states when compared with
In fact, salinity controls the behavior of F1 in a Crude-Brine-Rock (CBR)
viscous forces. It is evident that the reduction of capillary pressure
system in two ways, effecting mobility ratio and shifting end-point sat­
would contribute to the increase of breakthrough recovery factor
urations. To investigate the effect of this dimensionless number on the
(Fig. 14). That is, the lower the difference between two states of high
results, three different functions were used (Table 6). For the sake of
salinity and low salinity, the higher would be the breakthrough recovery
clarity, these functions are plotted versus dimensionless salinity
factor.
(Fig. 12).
Fig. 13 shows the effect of F1 on the propensity of BRF versus Ng. As it
ðPHS LS

can be seen, BRF increases with this dimensionless number. In this case, 3.2.10. Effect of (Npc ¼ Cow PCow Þ
PHS
) on breakthrough recovery factor
Cow
at initial condition, MHS was over one, which implies inappropriate The other dimensionless group related to the capillary pressure is
Npc. This number shows the effect of capillary pressure during LSWF in
LS
mobility ratio, followed by faster fingering flow. On the contrary, MwHS
LS relation to the initial capillary pressure. In other words, the difference
and Mo alter system towards appropriate mobility ratios, resulting in
HS
between capillary pressures is not the only criterion. As it can be seen
postponing breakthrough time. In turn, as F1 increases the tendency of
from Fig. 15, a decrease in capillary and gravity numbers results in
the system towards lower mobility ratio would also be higher.
higher breakthrough recovery factor. In addition, increase in BRF ensues

10
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

LS
Fig. 7. Breakthrough recovery factor versus Ng for different MoHS .

LS
Fig. 8. Breakthrough recovery factor versus Ng for different MwHS .

from the increase in capillary pressure ratio under similar conditions. Novakovic, 2002 also examined the impact of Nρ and similar trend
Aside from the effects of these two numbers on increase of break­ was observed.
through recovery factor, at higher capillary and gravity numbers, this � Gravity number (Ng ) was important at field scale but not a crucial
increase is steeper than the lower ones. This reaffirms the role of viscous factor at core scale.
forces in effectiveness of this dimensionless number. LS
� Mobility ratio scaling groups (MHS ;MoHS and MHS
LS
HS
w ) especially M had a
After all sensitivity analyses, the importance of each scaling group profound impact on reservoir scale results. This implies that fluids at
was characterized upon which it can be drawn that: core scale should have the similar mobility as reservoir state.
� RL controls VE in which the effects of gravity forces are influenced by
It was shown that Nρ was not an important factor at field scale. It this number.
should be noted that the density ratio was changed while the density � Salinity number ranges (CSD ) would be the controlling factor for
difference kept constant. Therefore, at core scale neglecting this initial and final states of salinity in the porous media. Therefore, it is
number would not affect the results of field scale. The effect of imperative to be considered at the reservoir conditions.
density difference (gravity segregation) would though be dominant � Nα can reverse the impact of gravity forces and its higher values
and reflected in Ng . Note that this conclusion is compliance with would result in higher BRF when Ng increases. However, in small
outcomes of previous studies. For instance, Shook et al. (1992) and gravity number no change is observed.

11
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Fig. 9. Breakthrough recovery factor versus Ng for different MHS .

Fig. 10. Breakthrough recovery factor versus Ng for different CInj


SD .

� F1 as interpolation parameter which controls the shifting between and inspectional analyses. The following conclusions can be drawn from
two states of oil-wet and water-wet. Generally, it is considered as the results of this study:
linear function of salinity, albeit deviation from this linearity can
alter BRF. � Ten scaling groups need to be used to scale LSWF from core scale to
� Capillary scaling groups (NC and Npc ) as representative of capillary field scale. However, for salinity scaling group, two extremes of this
forces can alter the relationships between gravity and viscous forces. number from high to low salinity relative permeabilities should be
Thus, they should be considered because of their influence at the same in the reservoir model.
different scales. � Similar to waterflooding approach, for LSWF, it was observed that
when RL increases above a certain limit then the change of break­
4. Conclusions through recovery factor versus gravity number would only be mar­
ginal which would indicate a sign of VE (RL > 5).
LSWF scale up would help to make decision whether to use it as an � Density ratio seems to have negligible impact on breakthrough re­
EOR approach for field scale based on the core scale experiments. This covery factor in both tilted and non-tilted reservoirs. Thus, using ten
has thoroughly been investigated in this study using both dimensional

12
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Fig. 11. Breakthrough recovery factor versus Ng for different CInit


SD .

� It would be recommended to find the number of scaling groups based


Table 6
on DA, while IA is more appropriate for finding of the physically
Different functions of F1.
meaningful scaling groups.
Scenario No. Function Area

1 F1 ¼ ð1 CsD Þ 2 0.33 CRediT authorship contribution statement


2 F1 ¼ ð1 CsD Þ 0.50
3
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
F1 ¼ 1 CsD 0.67 Sabber Khandoozi: Data curation, Writing - original draft, Investi­
gation, Software. Mohammad Reza Malayeri: Supervision, Writing -
review & editing. Masoud Riazi: Supervision, Conceptualization,
scaling groups would be enough for scaling processes. Other pro­ Project administration, Writing - review & editing. Mojtaba Ghaedi:
posed scaling groups influence the breakthrough recovery factor. Methodology, Software, Writing - review & editing.
� During LSWF, scaling groups related to mobility ratio, capillary
forces, gravity segregation and VE dominate the flow behavior in a
vertical 2-D system with homogeneous permeability.

Fig. 12. F1 versus CsD for different functions of F1.

13
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Fig. 13. Breakthrough recovery factor versus Ng for different F1.

Fig. 14. Breakthrough recovery factor versus NC for different Npc.

Abbreviations

Abbreviation Definition
LSWF Low Salinity Water Flooding
EOR Enhanced Oil Recovery
DA Dimensional Analysis
IA Inspectional Analysis
BRF Breakthrough Recovery Factor
VE Vertical Equilibrium
2-D Two Dimensional
CBR Crude-Brine-Rock

Symbols and notations


Symbol Definition
L Length

14
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Fig. 15. Breakthrough recovery factor versus Npc for different NC and Ng.

H Height
∅ Porosity
kx or z Horizontal or vertical permeability
Δρow Density difference between oil and water
ρw Water density
ρo Oil density
g Gravitational constant
α 2D plane angle with horizontal plane
uT Injection velocity
uwx or ox x-direction velocity (water or oil)
uwz or oz z-direction velocity (water or oil)
t time
CLS
s Salinity of low salinity water
CHS
s Salinity of high salinity water
Cs Salinity of water
F1 or Gint Interpolation parameter
FLS
1 or F 1
HS
Interpolation parameter (Low or high salinity system)
krw Water relative permeability
kLS HS
rw or krw Relative permeability of water (Low or high salinity system)
LS HS
krw or krw Average relative permeability of water (Low or high salinity system)
kro Oil relative permeability
kLS HS
ro or kro Relative permeability of oil (Low or high salinity system)
LS HS
kro or kro Average relative permeability of oil (Low or high salinity system)
Swco Connate water saturation
SLS HS
wco or Swco Connate water saturation (Low or high salinity system)
Swmax Maximum water saturation
SLS
wmax or SHS
wmax Maximum water saturation (Low or high salinity system)
Pcow Oil-water capillary pressure
PLS HS
cow or Pcow Oil-water capillary pressure (Low or high salinity system)
LS HS
λw or λw Average water mobility (Low or high salinity system)
LS HS
λo or λo Average oil mobility (Low or high salinity system)
LS HS
Pc or Pc Average capillary pressure (Low or high salinity system)
LS HS
kro or kro Average oil permeability (Low or high salinity system)
LS HS
krw or krw Average water permeability (Low or high salinity system)
Po Oil pressure
Pw Water pressure

15
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

XYD ðe:g:SwD Þ Dimensionless Parameter (e.g. dimensionless saturation)


μw or o Water or oil viscosity
Gi ith dependent or independent group
aw ; CLðHÞS
w ; aCLðHÞS
o Parameters used in capillary pressure correlation, Skjaeveland et al. (1998)
kLðHÞS LðHÞS LðHÞS LðHÞS
ro;ww ; krw;ww ; kro;ow ; krw;ow ; cwd ; cod Parameters used in relative permeability correlation, Skjaeveland et al. (1998)

Appendix A

In this appendix, the governing equations are discussed using the procedure of non-dimensionalization. Initially, scaling procedure was applied for
transformation from dimensional to dimensionless space i.e. for a typical parameter X. The transformations were conducted as such that X ¼ X*1 XD þ X*2
similar to what was performed by Shook et al. (1992) for waterflooding. The transformed equations were as follows:
∂Sw ∂Cs ∂uwx ∂uwz
φCs þ φSw þ Cs þ Cs ¼0
∂t ∂t ∂x ∂z
∂Sw ∂Cs ∂F1 ∂uwx ∂uwz
→φCs þ φSw þ Cs þ Cs ¼0
∂t ∂F1 ∂t ∂x ∂z
� �
φ C*s1 CsD þ C*s2 S*w1 ∂SwD φ S*w1 SwD þ S*w2 C*s1 ∂CsD ∂F1
→ *
þ *
t1 ∂tD t1 ∂F1 ∂tD
� �
C* CsD þ C*s2 u*wx1 ∂uwxD C* CsD þ C*s2 u*wz1 ∂uwzD (26)
þ s1 þ s1 ¼0
x*1 ∂xD z*1 ∂zD
� � � �
C* ∂SwD S*w2 ∂CsD ∂F1
→ CsD þ s2 þ S þ
C*s1 S*w1
wD
∂tD ∂F1 ∂tD
� � � �
C* * * C*s2 * *
CsD þ s2 u t C þ u t
C*s1 wx1 1 ∂uwxD C*s1 wz1 1 ∂uwzD
sD
þ þ ¼0
x*1 S*w1 φ ∂xD z*1 S*w1 φ ∂zD

φS*w1 ∂SwD u*wx1 ∂uwxD u*wz1 ∂uwzD


þ * þ * ¼0
t*1 ∂tD x1 ∂xD z1 ∂zD
(27)
∂SwD t*1 u*wx1 ∂uwxD t*1 u*wz1 ∂uwzD
→ þ * * þ * * ¼0
∂tD φSw1 x1 ∂xD φSw1 z1 ∂zD

∂ðuwx þ uox Þ ∂ðuwz þ uoz Þ


þ ¼0
∂x ∂z
� �
1 ∂ u*wx1 uwxD þ u*ox1 uoxD 1 ∂ u*wz1 uwzD þ u*oz1 uozD
→ * þ * ¼0
x1 ∂xD z1 ∂zD
(28)
u* ∂ðuwxD þ uoxD Þ u*wz1 ∂ðuwzD þ uozD Þ
→ wx1 þ * ¼0
x*1 ∂xD z1 ∂zD

∂ðuwxD þ uoxD Þ x*1 u*wz1 ∂ðuwzD þ uozD Þ


→ þ * * ¼0
∂xD z1 uwx1 ∂zD
� �� �
k P* krw P*w1 ∂PwD x*1 ρw g sin α u*wx2
LS HS

uwx ¼ kx F1 * rw *w1 þ ð1 F1 Þ þ (29–1)


uwx1 x1 μw u*wx1 x*1 μw ∂xD P*w1 u*wx1
� �� �
k P* kro P*o1 ∂PoD x*1 ρo g sin α u*ox2
LS HS

uox ¼ kx F1 *ro *o1 þ ð1 F1 Þ þ (29–2)


uox1 x1 μo u*ox1 x*1 μo ∂xD P*o1 u*ox1
� �� �
k P* krw P*w1 ∂PwD z*1 ρw g cos α u*wz2
LS HS

uwz ¼ kz F1 *rw *w1 þ ð1 F1 Þ þ (29–3)


uwz1 z1 μw u*wz1 z*1 μw ∂zD P*w1 u*wz1
� �� �
k P* kro P*o1 ∂PoD z*1 ρo g cos α u*oz2
LS HS

uoz ¼ kz F1 *ro *o1 þ ð1 F1 Þ þ (29–4)


uoz1 z1 μo u*oz1 z*1 μo ∂zD P*o1 u*oz1
3 2 3
Z ðH z*1 Þ=z*1

z* u*wx1 uwxD þ u*wx2 dzD 7 6H z
*
7
ρ ¼ ρo þ Δ ρ 1 5g cos α4 * 2 zD 5 at x ¼ L; 8t (30)
H 0 uT z1

16
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

P*w1 PwD þ P*w2 ¼ P*o1 PoD þ P*o2


2 3
ðH Zz*1 Þ=z*1 �
u*wx1 uwxD þ u*wx2 dzD 7
� � (31)
6 z* H z*2
¼ Pwf þ 4ρo þ Δρ 1 5g cos α zD
H uT z*1
0

For the water phase:


P*w1 PwD þ P*w2 ¼ Pwf þ
2 3
ðH Zz*1 Þ=z*1 � � �
6 z *
u*wx1 uwxD þ u*wx2 dzD 7 H z*2
4ρo þ Δρ 1 5g cos α zD
H uT z*1
0

Pwf P*w2
�*
z1 ρo g cos α (32)
→PwD ¼ þ þ
P*w1 P*w1
3 2 3
ðH Zz*1 Þ=z*1 � �
z*1 Δρg cos α * * u*wx2 7 6H z
*
7
z1 uwx1 uwxD þ * dzD 5 � 4 * 2 zD 5
P*w1 uwx1 z1
0

For the oil phase:


P*o1 PoD þ P*o2 ¼ Pwf þ
2 3
ðH Zz*1 Þ=z*1 � � �
6 z*1 u*wx1 uwxD þ u*wx2 dzD 7 H z*2
4ρo þ Δρ 5g cos α zD
H uT z*1
0

Pwf P*o2
�*
z1 ρo g cos α (33)
→PoD ¼ þ þ
P*o1 P*o1
3 2 3
ðH Zz*1 Þ=z*1 � �
z*1 Δρg cos α * * u*wx2 7 6H z
*
7
z1 uwx1 uwxD þ * dzD 5 � 4 * 2 zD 5
P*o1 uwx1 z1
0

Po ¼ Pw þ PC

→P*o1 PoD þ P*o2 ¼ P*w1 PwD þ P*w2 þ F1 PCow þ ð1


LS HS
F1 ÞPCow
(34)
P* PwD P* P*
LS HS
F1 PCow þ ð1 F1 ÞPCow
→PoD ¼ w1 * þ w2 * o2 þ
Po1 Po1 P*o1

ðH Zz*2 Þ=z*1
z*1 �
u*wx1 uwxD þ u*wx2 dzD ¼ uT
H
0

ðH Zz*2 Þ=z*1 �
u*wx2

uT H (35)
→ uwxD þ *
dzD ¼ * *
uwx1 uwx1 z1
0

x*2 *
at xD ¼ ; 8t tD þ t*2
x*1 1

H ρo g cos α
PwD ¼ þ
P*w1
3 2 3
ðH Zz*2 Þ=z*1
Hkx krw Δρg cos αz*1
HS
7 6
uwxD dzD 5 � 4H z*1 zD
7
z*2 5 (36)
μw L
0

x*2 *
at xD ¼ ; 8t tD þ t*2
x*1 1

u*wz2 z*2 *
uwzD ¼ a tzD ¼ ; 8x xD þ x*2 ; t*1 tD þ t*2 (37)
u*wz1 z*1 1

17
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

u*wz2 H z*2 *
uwzD ¼ at zD ¼ ; 8x1 xD þ x*2 ; t*1 tD þ t*2 (38)
u*wz1 z*1

u*oz2 z*2 *
uozD ¼ at zD ¼ ; 8x xD þ x*2 ; t*1 tD þ t*2 (39)
u*oz1 z*1 1

u*oz2 H z*2 *
uozD ¼ *
at zD ¼ ; 8x1 xD þ x*2 ; t*1 tD þ t*2 (40)
uoz1 z*1

Using the similar procedure stated in Shook et al. (1992), some coefficients or terms were set to zero and one to maintain the primary form of
equations (Table 7). For instance, in equation (21), water vertical velocity in z-direction was equal to one. Therefore, the dimensionless term should be
considered as zero, resulting in u*wz2 ¼ 0.
Therefore, the remaining scale factors (parameters with asterisk) were estimated as per Equation (41):
S*w1 ¼ 1 Swmax Swco ðΔSÞ; S*w2 ¼ Swco
uT Lμw uT Lμo
P*w1 ¼ HS ; P*o1 ¼ HS ; P*w2 ¼ P*o2 ¼ Pwf
kx krw kx kro

x*1 ¼ L; z*1 ¼ H; x*2 ¼ z*2 ¼ 0


(41)
LΔSφ *
t*1 ¼ ; t2 ¼ 0
uT
uT H
u*wx1 ¼ u*ox1 ¼ uT ; u*wz1 ¼ u*oz1 ¼
L
u*wx2 ¼ u*wz2 ¼ u*ox2 ¼ u*oz2 ¼ 0

The following standard normalization process is proposed to attain dimensionless number while considering the effect of salinity in the process of
LSWF (Fig. 16). In fact, it is imperative in LSWF to have the identical salinity groups at both extremes (initial and injection conditions).
Cs CLS
CSD ¼ s
(42)
CHS
s CLSs

Table 7
Resultant transformation parameters.
No. Coefficient/term Equation number Value

1 u*wx1 t*1 26, 27 1


x*1 S*w1 φ
2 u*wz1 t*1 26, 27 1
z*1 S*w1 φ
3 x*1 u*wz1 28 1
z*1 u*wx1
4 HS
krw P*w1 29–1 1
u*wx1 x*1 μw
5 HS
kro P*o1 29–2 1
uox1 x*1 μo
*
6 u*wx2 29–1 0
u*wx1
7 u*ox2 29–2 0
u*ox1
8 u*wz2 29–3 0
u*wz1
9 u*oz2 29–4 0
u*oz1
10 Pwf P*w2 32 0
P*w1
11 P*w2 P*o2 33 0
P*o1
12 u*wx2 33 0
u*wx1
13 x*2 35, 36 0
x*1
14 x*2 37, 38, 39, 40 0
15 t*2 37, 38, 39, 40 0
16 z*2 37, 39 0
z*1
17 u*wz2 37, 39 0
u*wz1
18 u*oz2 38, 40 0
u*oz1

18
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Fig. 16. Schematic of salinity function, a linear relationship.

By obtaining the equivalent of scale factors, they can be rewritten as:


� � � �
CLS ∂SwD Swr ∂CsD ∂F1
CsD þ HS s LS þ SwD þ
Cs Cs ∂tD 1 Sor Swr ∂F1 ∂tD
� LS � � LS �
C ∂uwxD C ∂uwzD
þ CsD þ HS s LS þ CsD þ HS s LS ¼0
Cs Cs ∂xD Cs Cs ∂zD
(43)
∂SwD ∂CsD ∂F1
→ðCsD þ G1 Þ þ ðSwD þ G2 Þ
∂tD ∂F1 ∂tD
∂uwxD ∂uwzD
þðCsD þ G1 Þ þ ðCsD þ G1 Þ ¼0
∂xD ∂zD

∂SwD ∂uwxD ∂uwzD


þ þ ¼0
∂tD ∂xD ∂zD
(44)
∂SwD ∂uwxD ∂uwzD
→ þ þ ¼0
∂tD ∂xD ∂zD

∂ðuwxD þ uoxD Þ ∂ðuwzD þ uozD Þ


þ ¼0 (45)
∂xD ∂zD
LS
!� HS �
krw ∂PwD kx krw ρw g sin α
uwx ¼ F1 HS
þ ð1 F1 Þ þ
krw ∂xD μw uT
0 1 (46–1)
� �
B LS C ∂PwD
→uwx ¼ @F1 M w þ ð1
HS F1 ÞA þ G3
∂xD

� LS �� HS �
kro ∂PoD kx kro ρo g sin α
uox ¼ F1 HS þ ð1 F1 Þ þ
kro ∂xD μo uT
0 1 (46–2)
� �
B LS C ∂PoD
→uox ¼ @F1 M o þ ð1
HS F1 ÞA þ G4
∂xD

19
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

LS
!� HS �
kz L2 krw ∂PwD Hkx krw ρw g cos α
uwz ¼ F1 HS þ ð1 F1 Þ þ
kx H 2 krw ∂zD uT Lμw
0 1 (46–3)
� �
B LS C ∂PwD
→uwz ¼ R2L @F1 M wHS þ ð1 F1 ÞA þ G5
∂zD

� LS �� HS �
kz L2 kro ∂PoD Hkx kro ρo g cos α
uoz ¼ 2
F1 HS þ ð1 F1 Þ þ
kx H kro ∂zD uT Lμo
0 1 (46–4)
� �
B LS C ∂PoD
→uoz ¼ R2L @F1 M oHS þ ð1 F1 ÞA þ G6
∂zD

HS HS LS HS HS
krw PwD F1 kx kro PCow þ ð1 F1 Þkx kro PCow
PoD ¼ HS
þ
kro uT Lμo (47)
HS
→PoD ¼ M PwD þ F1 G7 þ G8
Z 1
uwxD dzD ¼ 1 at xD ¼ 0; 8t at xD ¼ 0; 8t (48)
0

� HS
Hkx krw ρo g cos α
PwD ¼ þ
LuT μw
3 2 3
HS Z1
Hkx krw Δρg cos α
uwxD dzD 5 � 41 zD 5
LuT μw
0 (49)
2 3
Z1
→PwD ¼ 4G9 þ G10 uwxD dzD 5 � ½1 zD �
0

at xD ¼ 0; 8t

Table 9
Table 8
Range of parameters which were used for sensitivity analyses of each scaling
Resultant dimensionless numbers using IA.
group.
No. Dimensionless Group Symbol Dimensionless Number
No. Target scaling group Target Range of changes
1 Gint F1 ¼ fðCSD Þ parameter(s)
Lower Upper
2 LS LS (Unit)
kro bound bound
MoHS HS
kro
1 Effective aspect ratio (RL ¼ kz ðmDÞ 1 50
3 LS LS
krw sffiffiffiffiffi
MHS HS L kz
w krw )
H kx
4 MHS HS
krw 2 L α ðdegreeÞ 0 70
Dip number (Nα ¼ tanα)
HS H
krosffiffiffiffiffi 3 ρ 69, 59 40, 30
5 RL Density ratio (Nρ ¼ w ) ρw ; ρo ðlb =cu:ftÞ
L kz Δρ
H kx 4 LS LS kLS 0.4 1.0
ro;ww ð Þ
kro
6 Nρ (MoHS ¼ HS
ρw
)
Δρ kro
7 G3 HS
kx krw ρw g sin α 5 LS kLS 0.3 0.7
LS rw;ww ð Þ
krw
μw uT (MwHS ¼ HS )
8 G4 HS
kx kro ρo g sin α krw
6 High salinity water mobility μw ðcPÞ 0.5 1.0
μo uT HS
9 G5 HS
Hkx krw ρw g cos α krw μo
ratio (MHS ¼ HS
)
uT Lμw kro μw
10 G6 HS
Hkx kro ρo g cos α 7 Injection salinity number CInj
s ðlb =STBÞ
0.35 10.00
uT Lμo CInj CLS
(CInj ¼ HS s s
)
11 G7 HS HS LS
kx kro ðPCow PCow Þ
SD
Cs CLS
s
8 Initial salinity number (CInit CInit 10 20
uT Lμo SD ¼ s ðlb =STBÞ
12 G8 HS HS
kx kro PCow CInit
s CLS
s
)
uT Lμo CHS
s CLS
s
13 G9 HS 9 F1 F1 ( ) 0.33 0.67
Hkx krw ρo g cos α
10 Modified capillary number CLS HS
w ; Cw ðpsiÞ
5, 30 30, 55
LuT μw
HS LS
14 G10 HS
Hkx krw Δρg cos α kx kHS
ro ðPCow PCow Þ
(NC ¼ )
LuT μw uT Lμo
15 G11 HS 11 HS
ðPCow PCow Þ
LS
CLS HS
w ; Cw ðpsiÞ;
5, 10, 30, 60,
Hkx kro ρo g cos α (Npc ¼ )
HS uT ðft =sÞ 6e-7 1e-5
uT Lμo PCow
16 G12 HS
Hkx kro Δρg cos α 12 HS
kx kro Δρgcosα uT ðft =sÞ 3.0e-8 3.0e-5
(Ng ¼ )
uT Lμo uT μo

20
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Table 10
Values of constant scaling groups in sensitivity analyses.
No. Target scaling group Constant Scaling Groups

1 2 3 4 5 6 7 8 9 10 11

1 Effective aspect ratio (RL ¼ Ng ¼ Nα ¼ Nρ ¼ 6:1 LS= LS= MHS ¼ inj


CSD ¼ Cinit F1 ¼ NC ¼ 0 Npc ¼ 0
sffiffiffiffiffi
0:01 10 0 Mo HS ¼ Mw HS ¼ SD ¼
0:5
L kz 4:1 0 1
) 1:2 0:8
H kx
2 L Ng ¼ RL ¼ Nρ ¼ 6:1 LS= LS= MHS ¼ inj
CSD ¼ Cinit F1 ¼ NC ¼ 0 Npc ¼ 0
Dip number (Nα ¼ tanα)
H 0:01 10 7:5 Mo HS ¼ Mw HS ¼ SD ¼
0:5
4:1 0 1
1:2 0:8
3 Density ratio (Nρ ¼
ρw
) Ng ¼ RL ¼ Nα ¼ LS= LS= MHS ¼ inj
CSD ¼ Cinit F1 ¼ NC ¼ 0 Npc ¼ 0
0:008 7:5 0; 37:22 Mo HS ¼ Mw HS ¼ SD ¼
0:5
Δρ 4:1 0 1
4:083 1:2 0:8

4 LS Ng ¼ RL ¼ Nρ ¼ 6:1 Nα ¼ 0 LS= MHS ¼ inj


CSD ¼ Cinit F1 ¼ NC ¼ 0 Npc ¼ 0
kLS Mw HS ¼ SD ¼
(MoHS ¼ HSro
) 0:01 10 7:5 4:1 0 1 0:5
kro 0:8
5 LS Ng ¼ RL ¼ Nρ ¼ 6:1 LS= Nα ¼ 0 MHS ¼ inj
CSD ¼ Cinit F1 ¼ NC ¼ 0 Npc ¼ 0
kLS Mo HS ¼ SD ¼
(MwHS ¼ HSrw
) 0:01 10 7:5 4:1 0 1 0:5
krw 10
6 High salinity water mobility Ng ¼ RL ¼ Nρ ¼ 6:1 LS= LS= Nα ¼ 0 inj
CSD ¼ Cinit F1 ¼ NC ¼ 0 Npc ¼ 0
0:01 10 7:5 Mo HS ¼ Mw HS ¼ SD ¼
0:5
kHS
rw μo 0 1
ratio (MHS ¼ ) 1:2 0:8
kHS
ro μw
7 Injection salinity number Ng ¼ RL ¼ Nρ ¼ 6:1 LS= LS= MHS ¼ Nα ¼ Cinit F1 ¼ NC ¼ 0 Npc ¼ 0
0:01 10 7:5 Mo HS ¼ Mw HS ¼ 0
SD ¼
0:5
Inj CInj
s CLS
s 4:1 1
(CSD ¼ HS ) 1:2 0:8
Cs CLS
s
8 Initial salinity number Ng ¼ RL ¼ Nρ ¼ 6:1 LS= LS= MHS ¼ Cinj Nα ¼ 0 F1 ¼ NC ¼ 0 Npc ¼ 0
0:01 10 7:5 Mo HS ¼ Mw HS ¼ SD ¼
0:5
CInit CLS 4:1 0
(CInit
SD ¼ sHS s
) 1:2 0:8
Cs CLS
s
9 F1 Ng ¼ RL ¼ Nρ ¼ 6:1 LS= LS= MHS ¼ inj
CSD ¼ Cinit Nα ¼ NC ¼ 0 Npc ¼ 0
0:01 10 7:5 Mo HS ¼ Mw HS ¼ SD ¼
0
4:1 0 1
1:2 0:8
10 Modified capillary number Ng ¼ 9:33 RL ¼ Nρ ¼ 6:1 LS= LS= MHS ¼ inj
CSD ¼ Cinit F1 ¼ Nα ¼ 0 Npc ¼
7:5 Mo HS ¼ Mw HS ¼ SD ¼
0:5 0:5 0:7
kx kHS HS
ro ðPCow PLS
Cow Þ
4:1 0 1
(NC ¼ ) 1:2 0:8
uT Lμo
11 ðPHS P LS Ng ¼ 1:9 RL ¼ Nρ ¼ 6:1 LS= LS= MHS ¼ inj
CSD ¼ Cinit F1 ¼ NC ¼ Nα ¼ 0
Mo HS ¼ Mw HS ¼
Þ SD ¼
(Npc ¼ Cow HS Cow )
PCow 6:5 7:5 4:1 0 1 0:5 0:8
1:2 0:8 2:8

� HS
Hkx kro ρo g cos α
PoD ¼ þ
uT Lμo
3 2 3
HS Z1
Hkx kro Δρg cos α
uwxD dzD 5 � 41 zD 5
uT Lμo
0 (50)
2 3
Z1
→PwD ¼ 4G11 þ G12 uwxD dzD 5 � ½1 zD �
0

at xD ¼ 0; 8t

uwzD ¼ 0 at zD ¼ 0; 8xD ; tD (51–1)

uwzD ¼ 0 at zD ¼ 1; 8xD ; tD (51–2)

uozD ¼ 0 at zD ¼ 0; 8xD ; tD (51–3)

uozD ¼ 0 at zD ¼ 1; 8xD ; tD (51–4)


In what follows, the proposed procedure to specify the independent scaling groups is presented. Beforehand, two proposed scaling groups (i.e. G1
and G2) were eliminated from this process since G1 and G2 can be derived from CsD and SwD, respectively. To ensure the independency of the scaling
groups, using logarithm of each one, a linear algebra can be achieved which is appropriate to figure out the rank of a matrix. Scaling groups gained
using IA are shown in Table 8.
To obtain the number of independent scaling groups, they were changed into logarithmic form. Their coefficients are shown in the following
matrix. The rank of this matrix is equal to 11. Therefore, 11 independent scaling groups would be required to scale this process. However, by
considering CsD as an independent scaling group, the number of scaling groups would then be increased to 12.

21
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

0 1
B C
B C
B � C
B C
B log λHS
ro
C
B C
B log λHS � C
B rw C
0 1B � C 0 1
00 B C
LS
1 0 1 0 0 0 0 0 0 0 0 0 0 0 B log λro C logΠ 1
B0
B 1 0 1 0 0 0 0 0 0 0 0 0 0 00 CB log λ �
CB LS C B logΠ C
C B 2
B 1 00 CB rw C B logΠ C
B 1 0 0 0 0 0 0 0 0 0 0 0 0 CB C B 3 C
B0 00 CB logðLÞ C B logΠ C
B 0 0 0 1 1 0:5 0:5 0 0 0 0 0 0 CB C B 4 C
B0 0 0 0 0 0 0 0 1 0 1 0 0 0 00 CB logðHÞ C B logΠ C C
B CB
B
C B 5
C B logΠ C
B0 1 0 0 0 0 1 0 1 0 0 1 0 1 00 C logðkx Þ C
B CB
B
C B
C B
6 C
B1 0 0 0 0 0 1 0 0 1 0 1 0 1 C
00 CB logðkz Þ C B logΠ 7 C
B B C B C
B0 1 0 0 1 1 1 0 1 0 0 0 1 1 00 C C B logΠ 8 C
B CB
B logðρw gÞ C¼B C
B1 0 0 0 1 1 1 0 0 1 0 0 1 1 00 C C B logΠ 9 C
B CB
B C B C
B1
B 0 0 0 1 0 1 0 0 0 0 0 0 1 00 CB logðρo gÞ
C C B logΠ 10 C
C
B1 B C B
B 0 0 0 1 0 1 0 0 0 0 0 0 1 10 C
CB logðΔρgÞ C B logΠ 11 C
C
B1 CB C B
B 0 0 0 1 1 1 0 0 1 0 0 1 1 00 CB logðsinαÞ C B logΠ 12 C
C
B1 B C B
B 0 0 0 1 1 1 0 0 0 1 0 1 1 00 C
CB C B logΠ 13 C
C
B1 B logðcos αÞ C B
B 0 0 0 1 1 1 0 0 1 0 0 1 1 00 C
CB C B logΠ 14 C
C
@1 AB C @
0 0 0 1 1 1 0 0 0 1 0 1 1 00 B logðuT Þ � C logΠ 15 A
B C
0 0 0 0 0 0 0 0 0 0 0 0 0 0 01 B log PHSCow PLS
Cow
C logΠ 16
B C
B log PHS � C
B Cow C
B C
B logðF Þ C
B 1 C
B C
B C
@ A

Appendix B

In this appendix, the range of parameters, used in sensitivity analysis of each scaling group (Table 9) and values of other constant dimensionless
scaling groups (Table 10) are presented.

Appendix C. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.petrol.2020.106956.

References Dang, C., Nghiem, L., Nguyen, N., Chen, Z., Nguyen, Q., 2016. Mechanistic modeling of
low salinity water flooding. J. Petrol. Sci. Eng. 146, 191–209. https://doi.org/
10.1016/J.PETROL.2016.04.024.
Al-Ibadi, H., Stephen, K., Mackay, E., 2018. Improved Numerical Stability and Upscaling
Davis, J.A., Jones, S.C., 1968. Displacement mechanisms of micellar solutions. J. Petrol.
of Low Salinity Water Flooding. In: SPE Asia Pacific Oil and Gas Conference and
Technol. 20, 1415–1428. https://doi.org/10.2118/1847-2-PA.
Exhibition. Society of Petroleum Engineers. https://doi.org/10.2118/192074-MS.
Engelberts, W.F., Klinkenberg, L.J., 1951. Laboratory Experiments on the Displacement
Al-Ibadi, H., Stephen, K., Mackay, E., 2018. An updated fractional flow model of low
of Oil by Water from Packs of Granular Material. WPC.
salinity water flooding with respect to the impact of salt diffusion. In: SPE Trinidad
Geertsma, J., Croes, G.A., Schwarz, N., 1956. Theory of dimensionally scaled models of
and Tobago Section Energy Resources Conference. Society of Petroleum Engineers.
petroleum reservoirs. Trans. AIME 207, 118–127.
https://doi.org/10.2118/191222-MS.
GeoQuest, S., 2010. ECLIPSE Reservoir Simulator, Manual and Technical Description.
Al-Shalabi, E.W., Sepehrnoori, K., 2017. Low salinity and engineered water injection for
Gharbi, R., Karkoub, M., Elkamel, A., 1996. Artificial neural network for the prediction of
sandstones and carbonate reservoirs.
immiscible flood performance. Energy Fuels 9, 894–900.
Alvarado, V., Manrique, E., 2010. Enhanced Oil Recovery : Field Planning and
Gharbi, R., Peters, E., Elkamel, A., 1998. Scaling miscible fluid displacements in porous
Development Strategies. Gulf Professional Pub./Elsevier.
media. Energy Fuels 12, 801–811. https://doi.org/10.1021/ef980020a.
Andersen, P., Standnes, D.C., Skjæveland, S.M., 2017. Waterflooding oil-saturated core
Gharbi, R.B.C., 2002. Dimensionally scaled miscible displacements in heterogeneous
samples - Analytical solutions for steady-state capillary end effects and correction of
permeable media. Transport Porous Media 48, 271–290. https://doi.org/10.1023/A:
residual saturation. Journal of Petroleum Science and Engineering 157, 364–379.
1015723329598.
https://doi.org/10.1016/j.petrol.2017.07.027.
Greenkorn, R.A., 1964. Flow models and scalinkg laws for flow through porous media.
Attar, A., Muggeridge, A.H., 2016. Evaluation of Mixing in Low Salinity Waterflooding.
Ind. Eng. Chem. 56, 32–37. https://doi.org/10.1021/ie50651a006.
In: SPE EOR Conference at Oil and Gas West Asia. Society of Petroleum Engineers.
Greenkorn, R.A., Johnson, C.R., Harding, R.E., 1965. Miscible displacement in a
https://doi.org/10.2118/179803-MS.
controlled natural system. J. Petrol. Technol. 17, 1329–1335. https://doi.org/
Austad, T., Strand, S., Madland, M., Puntervold, T., Korsnes, R., 2008. Seawater in chalk:
10.2118/1232-PA.
an EOR and compaction fluid. SPE Reservoir Eval. Eng. 11, 648–654. https://doi.
Gupta, R., Mohanty, K.K., 2011. Wettability alteration mechanism for oil recovery from
org/10.2118/118431-PA.
fractured carbonate rocks. Transport Porous Media 87, 635–652.
Bagci, S., Kok, M.V., Turksoy, U., 2001. Effect of brine composition on oil recovery by
Hirasaki, G.J., 1980. Scaling of non-equilibrium phenomena In surfactant flooding. In:
waterflooding. Petrol. Sci. Technol. 19, 359–372.
SPE/DOE Enhanced Oil Recovery Symposium. Society of Petroleum Engineers. htt
Bahadori, A., 2018. Fundamentals of enhanced oil and gas recovery from conventional
ps://doi.org/10.2118/8841-MS.
and unconventional reservoirs.
Jadhawar, P.S., Sarma, H.K., 2008. Scaling and sensitivity analysis of gas-oil gravity
Berawala, D.S., Andersen, P.Ø., Ursin, J.R., 2019. Controlling parameters during
drainage EOR. In: Proceedings of SPE Asia Pacific Oil and Gas Conference and
continuum flow in shale-gas production: a fracture/matrix-modeling approach. SPE
Exhibition, pp. 20–22. https://doi.org/10.2118/115065-MS.
J. 24, 1378–1394. https://doi.org/10.2118/190843-pa.
Jerauld, G.R., Lin, C.Y., 2008. Webb., KJ, and Seccombe, JC, 2008. Modeling low salinity
British Petroleum, 2016. BP Energy Outlook 2016 98. https://doi.org/10.1017/
waterflooding. SPE Reservoir Eval. Eng. 11, 1000–1012.
CBO9781107415324.004.
Khorsandi, S., Qiao, C., Johns, R.T., 2017. Displacement efficiency for low-salinity
Carpenter, Bail, P.T., Bobek, J.E., 1962. A verification of waterflood scaling in
polymer flooding including wettability alteration. SPE J. 22, 417–430. https://doi.
heterogeneous communicating flow models. Soc. Petrol. Eng. J. 2, 9–12. https://doi.
org/10.2118/179695-PA.
org/10.2118/171-PA.
Korrani, A.K.N., Jerauld, G.R., 2018. Modeling wettability change in sandstones and
Craig, F., Sanderlin, J., Moore, D., Geffen, T., 1957. A laboratory study of gravity
carbonates using a surface-complexation-based method. In: SPE Improved Oil
segregation in frontal drives. Aime 210, 275–282.

22
S. Khandoozi et al. Journal of Petroleum Science and Engineering 189 (2020) 106956

Recovery Conference. Society of Petroleum Engineers. https://doi.org/10.2118/ Skjaeveland, S.M., Siqveland, L.M., Kjosavik, A., Hammervold, W.L., Virnovsky, G.A.,
190236-MS. 1998. Capillary pressure correlation for mixed-wet reservoirs. In: SPE India Oil and
Kwak, H.T., Zhang, G., Chen, S., 2005. The effects of salt type and salinity on formation Gas Conference and Exhibition. Society of Petroleum Engineers. https://doi.
water viscosity and NMR responses. In: Proceedings of the International Symposium org/10.2118/39497-MS.
of the Society of Core Analysts, Toronto, Canada, pp. 21–25. Sorbie, K.S., Feghi, F., Pickup, G.E., Ringrose, P.S., Jensen, J.L., 1994. Flow regimes in
Lake, L.W., 1989. Enhanced Oil Recovery. miscible displacements in heterogeneous correlated random fields. SPE Adv.
Lake, L.W., Johns, R., Rossen, B., 2014. Fundamentals of Enhanced Oil Recovery. Technol. 2, 78–87.
Leverett, M.C., Lewis, W.B., True, M.E., 1942. Dimensional-model studies of oil-field Sorbie, K., Mackay, E., 2000. Mixing of injected, connate and aquifer brines in
behavior. Trans. AIME 146, 175–193. waterflooding and its relevance to oilfield scaling. J. Petrol. Sci. Eng. 27, 85–106.
McGuire, P.L., Chatham, J.R., Paskvan, F.K., Sommer, D.M., Carini, F.H., 2005. Low https://doi.org/10.1016/S0920-4105(00)00050-4.
salinity oil recovery: an exciting new EOR opportunity for Alaska’s North Slope. In: Thomas, S., Ali, S.M., Thomas, N.H., 2000. Scale-up methods for micellar flooding and
SPE Western Regional Meeting. Society of Petroleum Engineers. their verification. J. Can. Petrol. Technol. 39.
Novakovic, D., 2002. Numerical Reservoir Characterization Using Dimensionless Scale van Daalen, F., van Domselaar, H.R., 1972. Scaled fluid-flow models with geometry
Numbers with Application in Upscaling. LSU Doctoral Dissertations. Louisiana State differing from that of prototype. Soc. Petrol. Eng. J. 12, 220–228. https://doi.org/
University, pp. 1–138. 10.2118/3359-PA.
Nygård, J.I., Andersen, P., 2019. Simulation of immiscible wag injection in a stratified Webb, K.J., Black, C.J.J., Edmonds, I.J., 2005. Low salinity oil recovery–The role of
reservoir - characterization of WAG performance. In: IOR 2019 - 20th European reservoir condition corefloods. In: IOR 2005-13th European Symposium on
Symposium on Improved Oil Recovery. European Association of Geoscientists and Improved Oil Recovery.
Engineers. EAGE. https://doi.org/10.3997/2214-4609.201900170. Yousef, A.A., Al-Saleh, S., Al-Kaabi, A., Al-Jawfi, M., 2011. Laboratory investigation of
Osif, T.L., 1988. The effects of salt, gas, temperature, and pressure on the compressibility the impact of injection-water salinity and ionic content on oil recovery from
of water. SPE Reservoir Eng. 3, 175–181. carbonate reservoirs. SPE Reservoir Eval. Eng. 14, 578–593. https://doi.org/
Parsons, R.W., 1977. Linear Scaling in slug-type processes application to micellar 10.2118/137634-PA.
flooding. Soc. Petrol. Eng. J. 17, 11–26. https://doi.org/10.2118/5846-PA. Zapata, V.J., Lake, L.W., 1981. A theoretical analysis of viscous crossflow. Society of
Patil, S.B., Dandekar, A.Y., Patil, S., Khataniar, S., 2008. Low salinity brine injection for Petroleum Engineers. https://doi.org/10.2118/10111-MS.
EOR on Alaska north slope (ANS). In: International Petroleum Technology Zeinijahromi, A., Nguyen, T.K.P., Bedrikovetsky, P., 2013. Mathematical model for fines-
Conference. https://doi.org/10.2523/IPTC-12004-MS. migration-assisted waterflooding with induced formation damage. SPE J. 18,
Perkins Jr., F.M., Collins, R.E., 1960. Scaling laws for laboratory flow models of oil 518–533. https://doi.org/10.2118/144009-PA.
reservoirs. J. Petrol. Technol. 12, 69–71. https://doi.org/10.2118/1487-G. Zhang, P., Tweheyo, M.T., Austad, T., 2007. Wettability alteration and improved oil
Pujol, L., Boberg, T.C., 1972. Scaling accuracy of laboratory steam flooding models. SPE recovery by spontaneous imbibition of seawater into chalk: impact of the potential
California Regional Meeting. Society of Petroleum Engineers. https://doi.org/10.211 determining ions Ca2þ, Mg2þ, and SO42 . Colloid. Surface. Physicochem. Eng.
8/4191-MS. Aspect. 301, 199–208.
Schwarz, N., Croes, G., 1955. Dimensionally scaled experiments and the theories on the Zhou, D., Yang, D., 2017. Scaling criteria for waterflooding and immiscible CO2 flooding
water-drive process. J. Petrol. Technol. 204, 35–42. in heavy oil reservoirs. J. Energy Resour. Technol. 139, 22909.
Seccombe, J., Lager, A., Jerauld, G., Jhaveri, B., Buikema, T., Bassler, S., Denis, J., Zohuri, B., 2015. Dimensional Analysis and Self-Similarity Methods for Engineers and
Webb, K., Cockin, A., Fueg, E., 2010. Demonstration of low-salinity EOR at interwell Scientists, Dimensional Analysis and Self-Similarity Methods for Engineers and
scale, Endicott field, Alaska. In: SPE Improved Oil Recovery Symposium. Society of Scientists. https://doi.org/10.1007/978-3-319-13476-5.
Petroleum Engineers. Al-Shalabi, E.W., Sepehrnoori, K., Pope, G., 2014. Mysteries behind the Low Salinity
Shook, M., Li, D., Lake, L.W., 1992. Scaling Immiscible Flow through Permeable Media Water Injection Technique. Journal of Petroleum Engineering 2014, 1–11. https://
by Inspectional Analysis, vol. 16. SITU-NEW YORK-, p. 311. doi.org/10.1155/2014/304312.

23

You might also like