Sei sulla pagina 1di 24

SPE-179695-MS

Displacement Efficiency for Low Salinity Polymer Flooding Including


Wettability Alteration
Saeid Khorsandi, Changhe Qiao, and Russell T. Johns, Pennsylvania State University

Copyright 2016, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Improved Oil Recovery Conference held in Tulsa, Oklahoma, USA, 11–13 April 2016.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Polymer flooding can significantly improve sweep and delay breakthrough of injected water, thereby
increasing oil recovery. Polymer viscosity degrades in reservoirs with high salinity brines, so it is
advantageous to inject low salinity water as a preflush. Low salinity water flooding (LSW) can also
improve local displacement efficiency by changing the wettability of the reservoir rock from oil wet to
more water wet. The mechanism for wettability alteration for low salinity waterflooding in sandstones is
not very well understood, however experiments and field studies strongly support that cation exchange
(CE) reactions are the key element in wettability alteration. The complex coupled effects of CE reactions,
polymer properties, and multiphase flow and transport has not been explained to date.
This paper presents the first analytical solutions for the coupled synergistic behavior of low salinity
waterflooding and polymer flooding considering cation exchange reactions, wettability alteration, adsorp-
tion, inaccessible pore volume (IPV), and salinity effects on polymer viscosity. A mechanistic approach
that includes the cation exchange of Ca2⫹, Mg2⫹ and Na⫹ is used to model the wettability alteration. The
aqueous phase viscosity is a function of polymer and salt concentrations. Then, the coupled multiphase
flow and reactive transport model is decoupled into three simpler sub-problems, one where cation
exchange reactions are solved, the second where a variable polymer concentration can be added to the
reaction path and the third where fractional flows can be mapped onto the fixed cation and polymer
concentration paths. The solutions are used to develop a front tracking algorithm, which can solve the slug
injection problem where low salinity water is injected as a preflush followed by polymer. The results are
verified with experimental data and PennSim, a general purpose compositional simulator.
The analytical solutions show that decoupling allows for estimation of key modeling parameters from
experimental data, without considering the chemical reactions. Recovery can be significantly enhanced by
a low salinity pre-flush prior to polymer injection. For the cases studied, the improved oil recovery (IOR)
for a chemically tuned LSP flood can be as much as 10% OOIP greater than with considering polymer
alone. The results show the structure of the solutions, and in particular the velocity of multiple shocks that
develop. These shocks can interact, changing recovery. For example, poor recoveries obtained in core
floods for small low salinity slug sizes are explained with intersection of shocks without considering
mixing. The solutions can also be used to benchmark numerical solutions and for experimental design. We
2 SPE-179695-MS

demonstrate the potential of LSP as a cheaper and more effective way for performing polymer flooding
when the reservoir wettability can be altered using chemically-tuned low salinity brine.
Introduction
Polymer flooding can significantly improve sweep efficiency and therefore enhance oil recovery (EOR)
(Sheng, 2010; Sheng et al., 2015). Combination of polymer flooding with other EOR methods such as gas,
alkaline and surfactant flooding has demonstrated synergistic effects that can lead to improved oil
recovery (Li et al. 2014; Luo et al. 2015; Sheng, 2014). The efficiency of polymer flooding greatly
depends on the salinity of the aqueous phase contacted (Sorbie 2013, Vermolen et al. 2011) because high
concentrations of monovalent and divalent ions reduce the polymer viscosity, and thus decrease the sweep
efficiency. In practice, a reservoir is pre-flushed using a low salinity water before polymer flooding to
avoid the mixing between the high salinity formation water and the polymer slug.
Recently, low salinity water flooding (LSW) is reported to improve oil recovery. In coreflooding
experiments, the chemical composition of the injection water is found to have a significant effect on oil
recovery. LSW can improve displacement efficiency by changing the wettability of a sandstone reservoir
from oil wet to more water wet (Morrow and Buckley 2011). Different mechanisms are proposed
including mineral dissolution, fine migration, surface potential change and multi-component ionic
exchange (MIE), among which the MIE mechanism (Lager et al. 2007) is the most supported with
experimental data and theoretical analysis (Sheng 2014b, Myint and Firoozabadi 2015). In this mecha-
nism, cation exchange between Na⫹, Ca2⫹ and Mg2⫹ is considered and how much Na⫹ is adsorbed by
the clay surface determines the wettability (Lager et al. 2007). As low salinity brine is injected, Na⫹ is
released from the surface and this process alters the surface affinity towards more water wet. The
wettability alteration leads to improved oil recovery, as measured in many coreflooding experiments
(Austad et al. 2012). Velderr et al. (2010) explained the connate water banking in field scale low salinity
floods as evidence for wettability alteration. The divalent cation concentrations usually reach levels below
the injection fluid, which indicate the presence of cation exchange reactions. Oil recovery steadily
increases in low salinity floods in sandstones even after many pore volumes of low salinity water injection
(RezaeiDoust et al. 2011, Shiran and Skauge 2013, Kozaki 2012), which indicates a slow moving
wettability alteration front.
Since low salinity water is used to pre-flush the reservoir for polymer, these two processes can work
together where wettability alteration and viscosity increase can improve both sweep efficiency and
microscopic displacement efficiency. Mohammadi and Jerauld (2012) developed a mechanistic model for
low salinity polymer flooding with relative permeabilities as a function of water salinity. Experimental
studies of the combined LSW and polymer flooding lead to very high oil recovery (Shaker Shiran and
Skauge 2013). However, to the best of our knowledge, there is not a conclusive mechanism that explains
how the low salinity and polymer interact with each other and what are the mechanisms that leads to such
a high recovery.
Significant advances have been made in recent years to predict wettability alteration and oil recovery
(Jerauld et al. 2008; Dang et al. 2013; Qiao et al. 2015a; Qiao et al. 2016). Jerauld et al. (2008) proposed
a fully compositional model that included the transport of salts in the aqueous phase as an additional
single-lumped component. They determined the relationship between the relative permeability and
residual oil saturation, but from linear interpolation of the wetting state based on total salinity without
tracking individual species. Korrani et al. (2014) coupled the UTCOMP reservoir simulator (Chang 1990)
and PHREEQC (Parkhurst and Appelo 1999) to model geochemical reactions. Dang et al. (2013)
developed a fully coupled geochemical and compositional flow model for low salinity waterflooding in
sandstones where cation exchange is believed to be the mechanism for improved oil recovery. However,
there is no discussion on how different species controls the processes. Qiao et al. (2015a) developed a
reservoir simulator for low salinity waterflooding in carbonates by considering geochemical reactions and
SPE-179695-MS 3

a mechanistic model for wettability alteration. There is currently a lack of a detailed representation of the
surface-geochemical reactions and the corresponding wettability alterations in multiphase-flow models,
and there is no simulation study that has considered both wettability alteration caused by cation exchange
reaction in low salinity waterflooding in sandstones and the increased viscosity of polymer.
Seccomb et al. (2008) coreflood experiments show a bank of low salinity water and no recovery for
a small slug of low salinity. They explained that the low recovery was due to dispersion of the small slugs,
but no consideration was given to the potential of interacting shocks since a mathematical model of this
process is lacking.
Analytical methods for EOR are based on the analytical solution of dispersion-free 1-D flow (Buckley
and Leveret 1942, Helfferich and Klein 1970, Helfferich 1981, Pope 1980, Dindoruk 1992, Johns 1992,
Dumore et al. 1984, Orr 2007, Lake et al. 2014). The 1-D convective displacements are modeled with
hyperbolic equations, which along with constant initial and injection conditions, are called Riemann
problems. The ultimate goal of solving Riemann problems is to find a global Riemann solver for a specific
type of problem. Once found, the Riemann solver can be used in front tracking methods (Holden and
Risebro 2013) that consider rarefactions as small shocks in order to solve for complex slug injections such
as water alternating gas floods and a polymer low salinity pre-flush. In addition the solvers can be used
in numerical simulators for higher order estimation of the fluxes. Such a solver and front tracking
algorithms are developed for polymer floods (Issacson 1989), polymer flood with adsorption (Johansen
and Winther 1989), first contact miscible WAG (Juanes and Lie 2008), two-component partially miscible
gas floods (Johns 1992), three-component partially miscible gas floods (Khorsandi et al. 2015) and
polymer injection with variable salinity (de Paula and Pires 2015, Borazjani et al. 2016).
Analytical solutions for cation exchange reactions have been developed for single phase transport.
Helfferich and Klein (1970) applied coherence theory to chromatographic separation. Pope et al. (1978)
constructed solutions for monovalent-divalent exchange. Appelo et al. (1993) calculated intermediate
concentrations for ion exchange transport by assuming that variations consist of shocks only. Venkatra-
man et al. (2014) has developed a Riemann solver for single phase transport with cation exchange
reactions.
The analytical solutions for complex LSP processes cannot be easily constructed using conventional
MOC. The splitting of the physical equations has shown great performance to simplify the problem. The
calculation of tie-line routes in gas floods independent of fractional flow has been studied by many
researchers (Johns 1992, Dindoruk 1992, Bedrikovetsky and Chumak 1992, Entov 1997). Pires et al.
(2006) expanded on this idea and developed Lagrangian coordinates to split the equations into two parts;
a set of equations dependent only on phase behavior, and one additional equation based on fractional flow.
Khorsandi et al. (2015) improved the splitting approach and published a global Riemann solver and front
tracking solutions for three component partially miscible gas floods. Khorsandi and Johns (2015) extended
the splitting technique to multicomponent gas displacements to separate the tie-line solution from
fractional flow and estimate the minimum miscibility pressure for many complex fluids such as bifur-
cating phase behavior (Ahmadi et al. 2011, Khorsandi et al. 2014) and real field data. de Paula and Pires
(2015) and Borazjani et al. (2016) used the splitting approach to develop a front tracking algorithm for
polymer injection with variable salinity. They did not consider, however, ion adsorption or wettability
alteration.
In this paper, we develop the first analytical solutions for the complex coupled process of low
salinity-polymer (LSP) slug injection in sandstones that identifies the key parameters that impact oil
recovery for LSP, and also improves our understanding of the synergistic process, where cation exchange
reactions change the surface wettability. Both secondary and tertiary LSP is considered. First, the paper
presents the mathematical numerical and analytical models. Then, splitting of the analytical equations are
developed along with analytical solutions for different scenarios. The developed analytical solutions are
4 SPE-179695-MS

validated against experimental results and numerical simulation. The velocity of the different ion and
saturation fronts are compared to demonstrate the insight of the new analytical solutions.

Mathematical model
We use the in-house general purpose compositional simulator, PennSim (PennSim 2013, Qiao 2015), to
make all numerical simulation calculations. The basic equations needed to model LSP flooding are
outlined here.
Mass conservation of oil, water, polymer and aqueous ionic species are included along with cation
exchange reactions, adsorption of salts and polymer, inaccessible pore volume, and wettability alteration.
A mechanistic approach that includes the cation exchange of Ca2⫹ and Na⫹ is used to model the
wettability alteration. The viscosity is a function of polymer and ionic species concentration.
Immiscible oil/water flow
The mass conservation equations for immiscible oil and water phases are as follows:
(1)

Darcy’s law governs the flow rate of each phase,


(2)

The subscript ⬙w⬙ refers to the water phase, while ⬙o⬙ to the oil phase. Capillary pressure relates the
pressure of oil and water phases,
(3)

The saturation relation completes the set of equations


(4)

The primary unknowns for the multiphase flow system are Po and Sw.
Cation Exchange Reaction Network
The cation exchange between clay and the aqueous phase is assumed to be the main mechanism for
wettability alteration. The primary cations include Na⫹, Ca2⫹ and Mg2⫹. With Na-X, Ca-X2 and Mg-X2
representing the surface sites occupied by sodium, calcium and magnesium, the cation exchange reactions
can be written as

where Keq,Ca and Keq,Mg are the reaction equilibrium constants for Ca2⫹ and Mg2⫹ exchange reactions,
respectively. Since the surface reactions occur very fast, it is usually assumed that the cation exchange
reactions are in equilibrium. For the above reactions, the mass action law is written as

where () represents thermodynamic activities. Here, for the convenience of the analytical solutions, we
assume dilute aqueous solution and the activities for surface species are
SPE-179695-MS 5

where [] denote the concentration in mol/g solid and CEC represents the total surface site concentration
in mol/g as

Reactive Transport Model


The mass conservation equations for the primary species p is
(5)

where the first term is the accumulation of total moles and the second term is the total molar flux of
the primary component p. The equations are based on the stoichiometric relationship among the species
participating in reactions. The derivation of the general reactive transport equations can be found in Qiao
et al. (2015b). For the cases considered in this paper, the reactive transport equations for Na⫹ Ca2⫹ and
Mg2⫹ are written as

Wettability alteration
We use a linear interpolation model as follows:
(6)

where , and are water relative permeabilities at the end-point oil-wet (ow) and end-point
water-wet states (ww). The same linear interpolation is used for other coefficients of the relative
permeability model. These end-point states do not have to be at the complete oil-wet or water-wet states,
but ideally should be measured at initial reservoir conditions (mixed wet state), and at the most water-wet
state possible (state achieved during LSW). The Brooks-Corey model is used (Brooks and Corey 1966),

where the normalized water saturation S* is calculated by


6 SPE-179695-MS

Here Swr is the residual water saturation and Sor is the residual oil saturation that depends on
wettability:

We further assume that the wettability alteration is controlled by the surface concentration of adsorbed
Na⫹, namely
(7)

Polymer Flooding Model


Polymer is dissolved and well mixed in the aqueous phase. The mass conservation equation for polymer
is
(8)

where ␾IPV is the inaccessible pore volume and Ĉp is the polymer that is adsorbed on the rock surface.
The viscosity of polymer solution is a function of polymer, Na⫹ and Ca2⫹ concentrations as (Delshad et
al. 1996)

The adsorbed polymer concentration Ĉp is a function of aqueous polymer concentration Cp, which is
determined from a table-lookup function. The shear rate dependence and viscoelastic effects of the
polymer are not considered. The residual oil saturation decreases during polymer flooding as a function
of the trapping number (Delshad and Pope, 1989).
Numerical solution
We used a finite volume method to discretize the PDEs. For each control volume k, the pressure Pj,k, water
saturation Sw,k and molar concentrations Ci,k are assumed to be at the geometric center. The volumetric
flow rate is evaluated at the interface between two control volumes using a central finite difference scheme
and upstream weighing. The temporal discretization uses a generalized non-iterative IMPEC solution,
which treats the pressure variable using the backward Euler method and the total moles of primary species
using the forward Euler method. A speciation calculation is performed after pressure and mole numbers
are calculated. The last step is to update the properties that include the effects of surface reactions on
porous media properties such as changing wettability. A more detailed solution procedure can be found
in Qiao (2015).
Model equations for analytical solutions
The mathematical model is simplified assuming 1-D incompressible dispersion-free flow. In addition,
reaction kinetics are ignored so that chemical reactions are always in equilibrium. Mass conservation of
oil, salt components and polymer (Eqs. 1, 5 and 7) are then given by,
(9)

(10)

(11)
SPE-179695-MS 7

where and fo ⫽ 1 – fw, xD, xD and tD are dimensionless distance and

time, and NC is the number of cations. The oil conservation equation can be rewritten using So ⫽ 1 – Sw
as
(12)

The conservation equations for polymer and ions can be expanded using the chain rule and simplified
using the above equation. That is,
(13)

(14)

where . In the next section, the analytical solutions are developed by splitting the flow equation
into three sub problems.

Decoupled system of equations


We define the new coordinates (Pires et al. 2006) as
(15)

(16)

The elements of Eqs. (12) are transformed therefore to the new coordinates as
(17)

The same calculation can be repeated for Eqs. (13) and (14). The final result of the transformation to
the new coordinates after some manipulation is
(18)

(19)

(20)

The hyperbolic equations (Eqs. 18 – 20) have the same form as the conservation equations when ␺ is
the time, ␸ is the location, fw⫺1, Cp and Ci are the conserved quantities, and Sw/fw, Ĉp, and are
the flux functions. Therefore we can use the method of characteristics (MOC) to solve the equations by
defining the characteristic velocity, ␴, in the new coordinates as
8 SPE-179695-MS

(21)

where is the eigenvalue of the characteristics matrix. Furthermore, represents the front retardation
for two-phase flow. Fronts with larger retardation appear later; hence, the solution is single valued when
decreases from the injection composition to initial composition. When ␺ ⫽ ⫺xD, the eigenvalues
increase from the injection composition to initial composition. We can convert the PDEs of Eqs. (18 - 20)
to ordinary differential equations using the definition of . The characteristics equations of Eqs. (18 - 20)
are
(22)

(23)

(24)

where . The following equations demonstrate the characteristic matrix for low salinity

polymer floods with two cations and one anion. Two independent ion concentrations are necessary to
calculate the equilibrium composition of the water phase and solid surface.
(25)

The eigenvalues of this system are


(26)

(27)

(28)

The system of equations is not strictly hyperbolic and the eigenvalues are not ordered based on their
values. The corresponding eigenvectors are as follows
(29)
SPE-179695-MS 9

where,

The eigenvectors and eigenvalues, Eqs. (26 – 29), have two important features. First, is the only
eigenvalue that is function of fractional flow and saturation. Second, composition is constant along ẽ4.
Therefore, we can solve the reaction and polymer transport independent of fractional flow and the path
along ẽ4. In addition, the polymer and reaction systems are uncoupled. The solution for the uncoupled
conservation equations can be constructed independently even when the system of equations are not
strictly hyperbolic. (For example uncoupled advection equations, Leveque 2002). Therefore, we first solve
for concentrations based solely on reaction and polymer transport. Then we map fractional flow onto these
concentrations. In this paper we assumed that adsorption of polymer is independent of salinity. In contrast,
the adsorption of polymer can be considered as a function of salinity, therefore the polymer transport
equation will be dependent on the reactive transport solution. Yet, the reactive transport solution will be
independent of polymer concentration.
The weak solution for Eqs. (18 - 20) in Lagrangian coordinates is equivalent to the weak solution of
the first equation (Wagner 1987). Therfore, we can calculate the shock velocities and determine the front
types based on the solution in Lagrangian coordinates.
Reactive transport
Equation (20) is similar to the single phase transport with cation exchange. Equation (30) is the
characteristic matrix for the single phase reactive transport with two cations (Venkatraman et al. 2014),
(30)

The eigenvalues for single-phase transport are related to the eigenvalues of Eqs. (24) as follows

such that a front with speed of 1.0 in the single-phase region has and , which implies no
retardation in the Lagrangian coordinates. However the eigenvectors of Eqs. (24) are the same as
single-phase reactive transport. The Riemann solver for the single phase reactive transport developed by
Venkatraman et al. (2014) can be used to construct solutions for low salinity cation exchange reactions
in low salinity flooding. The important features of the solutions of cation exchange reactions are as
follows. We assumed the anion is not adsorbed, therefore the anion front is a contact discontinuity with
characteristic speed of 1.0. Therefore, the retardation of the anion, , is always zero in Lagrangian
coordinates and the anion front moves ahead of the cation exchange front. In addition, as a result of the
constant CEC assumption, the anion shock has no effect on the surface composition. The cation fronts for
low salinity injection are always shocks and the front for high salinity injection are always rarefactions
with negligible retardation. Therefore, the anion shock has no effect on surface concentrations and as a
result wettability is not altered with the anion shock.
Polymer transport
The polymer eigenvalue, , is the retardation factor for the polymer front. The eigenvalue is calculated
based on the slope of the adsorption isotherm. As mentioned earlier, although the polymer viscosity is a
function of salinity, the MOC solutions for polymer in Lagrangian coordinates are independent of salinity
solutions. The salinity of water can affect the adsorption of the polymer. In that case, the polymer
concentration will change along with the reactive transport eigenvectors. However the reactive transport
10 SPE-179695-MS

system remains independent of polymer transport. The shock velocity can be calculated using the
Rankine-Hugoniot condition as

The adsorption isotherms are usually concave so that the solutions for polymer injection with polymer
adsorption always have a shock in the polymer concentration. In contrast, the solutions for injection of
chase fluid in a polymer flood exhibit a rarefaction wave for polymer concentration due to gradual
desorption of polymer from the rock surface. Polymer adsorption and porosity degradation is commonly
considered as an irreversible process, therefore the polymer will not desorb from the rock surface and the
solution for chase fluid injection will have a shock in the polymer concentration. Furthermore, non-
Newtonian behavior of polymer can be incorporated in the current solution (Rossen et al. 2010).
Map fractional flow
The last step of constructing the solution is to map fractional flow onto concentration solutions. Our
solution can be considered as an extension of fractional flow theory to multicomponent systems.
Alternatively, we can construct the complete solution in Lagrangian coordinates using Eqs. (22), then
transform the solution to xD – tD coordinates. The solution construction for compositional shocks are
sufficient to construct the low salinity polymer injection solution because rarefactions only occur for
injection of a slug of polymer and low salinity. The slug injection problem can be solved using front
tracking algorithms where rarefactions are estimated with several shocks. The shock velocity in xD – tD
coordinate is ⌳ ⫽ ⌬xD/⌬tD then . This means the extension of a C-shock should pass through
the point on the Sw axis.
Now, we demonstrate the steps to construct the solutions for multiple cases with only one C-shock and
. Figure 1 (left) demonstrates two hypothetical fractional flows for upstream and downstream
compositions of the shock. The shock between these two water composition states has . The
change in fractional flow properties is a result of a change in polymer concentration and/or wettability
alteration. The solutions for one upstream composition, L, and three downstream compositions, R1, R2 and
R3 are shown in Figure 1. The C-shock is independent of the downstream composition. The solutions
consist of a rarefaction form L to a, then a tangent shock to b. The final part of the solution can be
constructed as a Buckley leveret problem with b as injection composition and Ri as initial composition.
The profiles for these three problems are shown in Fig. 1 (right).

Figure 1—Mapping of fractional flow curve to the composition solution. Left figure uses the standard approach as is solved for the
fractional flow problem for polymer flooding. Right figure demonstrates the wave velocities for the three different Riemann problems.

Figure 2 shows the C-shock for different upstream compositions. The C-shock is always a function of
the downstream composition. For example the solution for L1 – R1 is a shock to a then a jump between
SPE-179695-MS 11

two fractional flows that is tangent at b and a rarefaction to R1. The second solution is a shock from L2
to c followed by a Buckley-Leverett shock to R1. Finally, the last composition path is a shock from L3 to
d followed by saturation shock to R2. The examples in Figs. 1 and 2 demonstrate that the C-shock is
always a function of the upstream composition; therefore, the solution for multiple C-shocks can be
constructed sequentially from the injection composition to initial composition without trial and error.

Figure 2—Mapping of fractional flow curve to composition solution. Left figure uses the same approach as fractional flow for polymer
flooding. Right figure demonstrate the wave velocities for the three different Riemann problems.

Front tracking algorithm


The analytical solutions for different combinations of injection and initial conditions are described in the
previous section. These solutions can be used to calculate the interaction of fronts for complex slug
injection problems and for varying initial conditions. The basic procedure is as follows. First, the
fractional flow is estimated with a piecewise linear function as shown in Figure 3. The solution for a
simple water flood based on the smooth piecewise fractional flow curves is shown in Figure 4. That is,
the rarefaction is converted to a series of shocks. The leading front velocity is slightly different between
the two solutions. The mass is, however, conserved in both cases. Furthermore, the accuracy of solution
can be increased by approximating the fractional flow curve with more linear pieces. The second step is
to estimate the initial and injection compositions with piecewise constant values. Figure 5 (left) shows
initial water saturation for a reservoir with stepwise initial water saturation. The front tracking algorithms
start by constructing solutions for the initial condition jumps at time zero. These jumps are shown by red
dots along the horizontal axis of Fig. 5 (right). Each line in Fig. 5 (right) represents a shock and saturations
have constant values between the lines. The shocks may intersect depending on their velocity. That is, the
upstream faster shock could catch up with the downstream slower shock as shown by point a in Fig. 5
(right). When they intersect, a new Riemann problem forms. The solution should be constructed for the
upstream saturation (Sw ⫽ 0.59) and downstream saturation (Sw ⫽ 0.1). The algorithm is finalized when
there are no additional shock intersections. More details of the front tacking algorithm can be found in
Holden and Risebro (2016).
12 SPE-179695-MS

Figure 3—Piecewise linear approximation of fractional flow is commonly used in front tracking algorithms.

Figure 4 —The piecewise estimate of the fractional flow curve converts the rarefactions to small shocks. The error of approximation
decreases as the number of the linear pieces of fractional flow is increased.

Figure 5—The interaction of shocks in a water flooding displacement with variable initial condition. The initial condition should be
approximated with a piecewise constant function.
SPE-179695-MS 13

Matching low salinity waterflooding experiments independent of reactions


We simplified the analytical solution by assuming that only one of the cation shocks alters the wettability,
which we call the wettability front. This assumption helps to reduce the cation exchange reaction model
to a single retardation coefficient for the wettability front. The retardation coefficient and the produced
water chemistry can be matched using a Riemann solver for single-phase reactive transport or by trial and
error using numerical simulators to calculate the reaction model parameters. The retardation coefficient
is a function of the low salinity water composition, CEC and reaction equilibrium coefficients. Although
the wettability front changes the surface composition significantly, the change in water composition is
smaller and the front is usually smeared out in the production data. Furthermore, when the high salinity
water is injected the fronts are rarefactions that move very fast, so that we can ignore the retardation effect
for high salinity injection in a low salinity flooded reservoir. Therefore the low salinity experiments can
be matched by these steps.
1. Measure viscosity, relative permeabilities, capillary pressures, polymer viscosities and polymer
adsorption isotherms for high and low salinity mixtures.
2. Match the recovery curves by adjusting the retardation coefficient for the wettability front.
3. Convert produced water composition to single phase data as described in next section. Then match
the compositions and retardation coefficient.

Match reactions independent of fractional flow


The single phase reactive transport simulation codes are commonly used to match the geochemical
reactions in low salinity floods. This estimation is valid because the oil saturation is usually very small
and close to residual saturation, therefore flow can be assumed to be single phase transport. We use the
splitting approach to eliminate the effect of fractional flow on experimental results. The Lagrangian
coordinate ␸ can be calculated using the recovery curves.
(31)

The parameter ␸ ⫹ 1 is the equivalent single phase flow time. Then we plot ion concentrations as a
function of ␸ ⫹ 1, which makes the results independent of fractional flow curve. The reactions, therefore,
can be matched independent of fractional flow.

Results
First we validate the analytical solutions for two-phase flow with cation exchange reactions, where
wettability alteration is initially neglected. Next, the effect of CE and wettability alteration on the
analytical solutions is demonstrated. Finally a low salinity polymer experiment and a series of low salinity
slug injection displacements are matched with the analytical solutions.

Two phase CEC without wettability alteration


In this case all eigenvalues are independent of each other and we have three completely decoupled
systems. The front velocities for salinities can be calculated as the slope of the line tangent to the curve
drawn from as shown in Figure 6. We used the injection and initial composition in Voegelin et
al. (2000) as shown in Table 1. Figure 7 demonstrates the analytical solutions and numerical simulation
results for the single- and two-phase transport with cation exchange reactions. The fronts for the two phase
case moves faster than the single-phase case because a portion of the pore volume is filled with oil. The
solution consists of an anion shock with zero retardation and two cation shocks, which are retarded for
7.5 and 16.3 PVI. The concentration of [Na – X]at xD ⫽ 1 is shown in Figure 8. The surface composition
is significantly changed by the first cation shock. Since the surface wettability is considered a function of
[Na – X], the first cation shock alters the surface wettability as it moves through the reservoir.
14 SPE-179695-MS

Figure 6 —The single phase CE reactions are converted to two-phase transport. The slope of dashed lines are equal to cation front
velocities.

Table 1—Water composition for the single- and two-phase displacements.


Voegelin et al. (2000) Shaker Shiran, and Skauge (2013) Seccombe et al. (2008)

Ion (mol/l) Injection Initial Injection Initial Injection Initial

Na⫹ 0.00166 0.47000 0.04940 0.49400 0.01001 0.45916


Mg⫹⫹ 0.00062 0.04900 0.00118 0.01178 0.00244 0.11175
Ca⫹⫹ 0.00310 0.01150 0.00547 0.05473 0.00043 0.01970
Cl⫺ 0.00910 0.59100 0.06270 0.92701 0.01574 0.72206

Figure 7—Comparison of single- and two-phase transport of Mgⴙⴙ. Wettability alteration is not included in this model. The simulation
results are shown with dotted lines.
SPE-179695-MS 15

Figure 8 —Comparison of single- and two-phase adsorbed concentration of Na at xD ⴝ 1 for the floods of Fig. 7. The surface
composition is not affected by the anion shock. The simulation results are shown with dotted lines.

Low salinity waterflooding


The analytical solutions for examples in the previous section are constructed with wettability alteration
caused by cation exchange reactions as shown in Fig. 9. The flow models are described in Table 2. The
sensitivity of the results to CEC is demonstrated in Fig. 9. Figure 10 demonstrates the analytical solutions
using fractional flow curves. The analytical solution using no CEC over predicts the wettability front
velocity.

Figure 9 —Comparison of analytical solution results (solid line) and simulation results (dotted line) for high salinity and low salinity
injection considering the effect of wettability alteration. The analytical solution with no CEC over predicts the effect of low salinity
injection.

Table 2—Reaction parameters. CEC2 and CEC3 are calculated by matching experiments.
CEC1 (mol/l)
Parameter Keq,Ca Keq,Mg (Venkatraman et al. 2014) CEC2 (mol/l) CEC3 (mol/l)

Value 46.933 67.8390 0.1171 3.1617 0.0033


16 SPE-179695-MS

Figure 10 —Solutions for low salinity water flooding. Left figure shows the analytical solution with original CEC and the right figure
shows the analytical solution without CEC. The wettability front velocity is over estimated in the right figure.

Low salinity polymer experiment


Shaker Shiran and Skauge (2013) low salinity polymer experiments are matched with our analytical
solutions. The input parameters for the model are shown in Table 2 and 3. The experiments are matched
by tuning the CEC, high salinity residual oil saturation, and the residual oil saturation reduction by
polymer. The matched CEC value is shown as CEC2 in Table 2. The experiments were conducted with
a low salinity slug followed by low salinity polymer buffer. The compositions of the water in both cases
are the same. We first demonstrate the analytical solution for low salinity polymer injection, then we used
the front tracking algorithm to match polymer slug injection.

Table 3—The Corey relative permeability parameters for the experiments.


Shaker Shiran, and Skauge (2013) Seccombe et al. (2008)

Parameters Water wet Oil wet Water wet Oil wet

Swr 0.22 0.22 0.15 0.07


Sor 0.10 0.16 0.24 0.29
nw 1.60 1.60 9.46 3.70
no 2.80 2.80 1.48 6.16
0.50 0.50 0.40 0.40
0.93 0.93 1.00 1.00
␮w 1.03 1.03 1.00 1.00
␮o 2.40 2.40 1.20 1.20
␮p 2.60 2.60 - -

The solution is presented using a Walsh diagram (Walsh and Lake 1989 and Lake et al. 2014) in Figure
11. The solution consists of a small rarefaction from J to a along the water-wet polymer fractional flow
curve (WWP) (Fig. 11 top, left) followed by a wettability front from a to b. Then, there is a rarefaction
along the oil-wet polymer curve (OWP) to c followed by a tangent polymer shock to d. The solution is
completed by a leading saturation shock along the oil-wet fractional flow curve (OW) from d to I. Figure
11 (top, right) shows recovery is poor because of the slow moving wettability front (shock ba). As shown
in Fig. 8, only the first cation exchange front changes surface wettability, therefore we only considered
one wettability front in the analytical solution of Fig. 11. The oil recovery is continued to 17 PVI for a
SPE-179695-MS 17

low salinity flood. The reaction model parameters are shown in Table 2. The cation exchange shock
velocity is matched by adjusting the CEC value shown as CEC2 in Table 2. Figure 12 demonstrates the
match between analytical solutions and experimental data for the low salinity polymer flooding experi-
ments by Shaker Shiran and Skauge (2013). Figure 13 gives a Walsh diagram for injection of a low
salinity water slug followed by polymer. The low salinity front moves very slow in the reservoir and the
polymer shock interacts with the low salinity shock even after a long period of low salinity injection as
shown in Fig. 13 (bottom right). The front tracking algorithm is used to calculate the analytical solution
after the polymer injection. A preflush of the reservoir with low salinity water is commonly used to
improve polymer flood performance. Figure 13 shows that the distance between polymer and high salinity
water increases in the reservoir even for a small slug of preflush. Therefore, the optimum low salinity
preflush can be determined based on the dispersion level in the reservoir.

Figure 11—Walsh diagram for low salinity polymer injection. Fractional flows are shown for oil wet (OW), oil wet with polymer (OWP)
and water wet with polymer (WWP). The anion and polymer shocks have the same velocity. The wettability front is very slow.
18 SPE-179695-MS

Figure 12—Analytical solution and simulation results matched experimental data (Shaker Shiran and Skauge 2013). CEC and oil wet Sor
were not provided for the experimental data and they are the only two fitting parameters used to match the low salinity flood.

Figure 13—Walsh diagram for low salinity flood followed by polymer injeciton. The fractional flows are shown for oil wet (OW), water
wet (WW), and water wet polymer (WWP). The solution is not self similar and the results are calculated by the front tracking algorithm.
SPE-179695-MS 19

Figure 14 —Low salinity pre-flush. The pink area shows the high salinity water and green area represents the polymer flooded region.

Low salinity slug injection with varying slug size


Seccombe et al. (2008) performed low salinity slug injection experiments with varying slug sizes. Their
results showed no oil recovery for small slugs, which was explained as the result of mixing. We used the
relative permeability data provided in the paper, and tuned the CEC to match their results with our
analytical solutions. The analytical solution is not affected by dispersion, but the oil is still not produced.
The analytical solutions demonstrate that the zero oil production for small slugs in the experiments is a
result of intersecting shocks, not dispersive mixing.
Figure 15 (left) demonstrates the fronts for a 0.2 pore volume low salinity slug. The high salinity slug
catches up to the wettability front at point b on Fig. 15 (left). Therefore no more oil is added to the oil
bank and the oil bank spreads in the core, significantly increasing the breakthrough time. Figure 15 (right)
demonstrates the fronts for 0.6 pore volume injection where the wettability front breaks through before
the high salinity chase water catches up, hence the oil bank is produced. Figure 16 shows the saturation
profiles at 15 PVI. The oil bank moves slowly and spreads out for 0.1 and 0.2 PVI injection. For the 0.3
PV low salinity slug experiment, the oil bank breaks through, but after a long time.

Figure 15—Saturation fronts for 1D low salinity slug injection. The low salinity slug size is 0.2 PV for left figure and 0.6 PV for the right
figure. The Naⴙ significantly reduces at the front shown by the red line so that wettability alteration occurs across this line. The shaded
region represents the water with very low salinity.
20 SPE-179695-MS

Figure 16 —Water saturation profiles for different low salinity slug sizes after 15 PVI calculated by MOC with cation exchange reaction.

The reduction in residual oil saturation is matched for different slug sizes by adjusting the CEC value
as shown in Figure 17. A relatively small CEC value (CEC3 in Table 2) was used to match the results,
and the simulation results were sensitive to dispersion so that a large number of grid blocks were required
to match the analytical solutions. The PennSim results using 100 grid blocks are shown in Figure 17.
Seccombe et al. (2008) concluded that the 0.2 PV low salinity slug is ineffective because of mixing,
however our analysis shows that the interaction of high salinity and wettability fronts can explain this
phenomena. The produced water chemistry is required to match the reaction parameter models more
precisely. Lager et al. (2011) examined the produced water geochemistry of the same reservoir and
concluded that the cation exchange reactions are possibly different from the aquifer freshening model
(Valocchi et al. 1981).

Figure 17—Comparison of Sor decrease from the analytical solutions to simulation and experimental results (Seccombe et al. 2008) for
different low salinity slug sizes after 15 PVI. The simulation model used 100 grid blocks.

Conclusions
Analytical solutions were constructed for low salinity polymer flood in sandstones considering a
mechanistic model of wettability alteration based on cation exchange reactions. The solutions were
SPE-179695-MS 21

developed by splitting the equations into reaction, polymer, and fractional flow parts. The solutions were
validated using numerical simulation and experimental data. The main conclusions are as follows.
1. The analytical solution results match the experimental data indicating that the proposed model for
wettability alteration through cation exchange is the likely mechanism.
2. The cation exchange front moves slower than the salinity shock (anion front). Surface composi-
tions and wettability alteration occurs only along the cation exchange front.
3. Oil recoveries are matched by adjusting parameters that retard the wettability front, without the
need to match the parameters for the cation exchange reaction model. After recoveries are
matched, the reaction model is tuned to match water composition, independent of fractional flow.
This makes the model parameters involved in tuning more reliable.
4. Oil is recovered gradually over several pore volumes of low salinity water injection. Most of the
oil is recovered once the wettability front breaks through.
5. Small slugs of low salinity water can be ineffective because the high salinity shock moves faster
than the wettability front, causing wettability alteration to cease at that point. The low salinity slug
should be of sufficient size to propagate the wettability front to the production wells.

Acknowledgement
The authors thank the member companies of the Enhanced Oil Recovery JIP in the EMS Energy Institute
at The Pennsylvania State University at University Park, PA for their financial support. Dr. Russell T.
Johns is Chair of the Petroleum and Natural Gas Engineering program and holds the Victor and Anna Mae
Beghini Faculty Fellowship in Petroleum and Natural Gas Engineering at The Pennsylvania State
University.

Nomenclature
() ⫽ thermodynamic activities of a species (dimensionless)
[] ⫽ the concentration of a solid species (mol/g solid)
Cs ⫽ the concentration of a species (mol/kg water or mol/g solid)
Ĉs ⫽ the adsorbed concentration of a species (mol/volume solid)
a1, a2, a3, sp ⫽ parameters used in viscosity model for polymer
Fp ⫽ molar rate of the primary species p (mol/m · day)
Fq ⫽ molar rate of the secondary species q (mol/m · day)
kra ⫽ relative permeability of phase ␣ (dimensionless)
⫽ endpoint relative permeability of phase ␣ (dimensionless)
⫽ water wet endpoint relative permeability of phase ␣ (dimensionless)
⫽ oil wet endpoint relative permeability of phase ␣ (dimensionless)
Keq,r ⫽ equilibrium constant of reaction r (dimensionless)
K ⫽ permeability (md or m2)
Mp ⫽ molar density of the primary species p (mol/m3)
Mq ⫽ molar density of the secondary species q (mol/m3)
Np ⫽ the number of the primary species
Nsec ⫽ the number of secondary reactions
n␣ ⫽ exponent in Corey’s model for phase ␣
Qp ⫽ total molar rate of primary species p (lbmol/day or mol/s)
S␣ ⫽ saturation of phase ␣ (dimensionless)
⫽ normalized water saturation of phase ␣ (dimensionless)
xD ⫽ dimensionless distance
22 SPE-179695-MS

tD ⫽ dimensionless time
⫽ velocity of phase ␣ (ft/day or m/day)
P␣ ⫽ pressure of phase ␣ (psi or Pa)
Pcow ⫽ capillary pressure between water phase and oil phase (psi or Pa)
␾IPV ⫽ inaccessible porosity (dimensionless) for polymer (dimensionless)
␾ ⫽ porosity (dimensionless)
␪ ⫽ wettability index (dimensionless)
␳␣ ⫽ density of phase ␣ (lb/ft3 or kg/m3)
␮␣ ⫽ viscosity of phase ␣ (cp)
vip ⫽ the (i, p) entry in the stoichiometry matrix for reactions in canonical form

References
Ahmadi, K., Johns, R.T., Mogensen, K., Noman, R., (2011). Limitations of current method-of-characteristics (MOC)
methods using shock-jump approximations to predict MMPs for complex gas/oil displacements, SPE Journal, pp.
743–750.
Appelo, C. A. J., Hendriks, J. A., Van Veldhuizen, M. (1993). Flushing factors and a sharp front solution for solute
transport with multicomponent ion exchange. Journal of Hydrology, 146, 89 –113.
Austad, T.; Shariatpanahi, S. F.; Strand, S.; Black, C. J. J.; Webb, K. J., (2012) Conditions for a Low-Salinity Enhanced
Oil Recovery (EOR) Effect in Carbonate Oil Reservoirs. Energy & Fuels, 26, (1), 569 –575.
Bedrikovetsky, P., Chumak, M., (1992). Riemann problem for two-phase four-and more component displacement (Ideal
Mixtures). In 3rd European Conference on the Mathematics of Oil Recovery.
Borazjani, S., Bedrikovetsky, P., Farajzadeh, R. (2016). Analytical solutions of oil displacement by a polymer slug with
varying salinity. Journal of Petroleum Science and Engineering, 140, 28 –40.
Brooks, R.H., Corey, A.T. (1966). Properties of porous media affecting fluid flow. Journal of the Irrigation and Drainage
Division, 92(2), 61–90.
Buckley, S.E. and Leverett, M.C., (1942). Mechanism of fluid displacement in sands, Trans., AIME, 146, pp. 107–116.
Chang, Y. B. (1990). Development and application of an equation of state compositional simulator.
Dang, C. T., Nghiem, L. X., Chen, Z., Nguyen, Q. P., & Nguyen, N. T. (2013). State-of-the art low salinity waterflooding
for enhanced oil recovery. In SPE Asia Pacific Oil and Gas Conference and Exhibition. Society of Petroleum
Engineers.
de Paula, A.S., Pires, A.P. (2015). Analytical solution for oil displacement by polymer slugs containing salt in porous
media. Journal of Petroleum Science and Engineering, 135, 323–335.
Delshad, M., Pope, G.A. (1989). Comparison of the three-phase oil relative permeability models. Transport in Porous
Media, 4(1), 59 –83.
Delshad, Mojdeh, G. A. Pope, K. Sepehrnoori. (1996). A compositional simulator for modeling surfactant enhanced
aquifer remediation, 1 formulation.⬙ Journal of Contaminant Hydrology 23.4, 303–327.
Dindoruk, B., 1992. Analytical theory of multiphase multicomponent displacement in porous media, Department of
petroleum engineering, Stanford, California, Stanford University.
Dumore, J.M., Hagoort, J., Risseeuw, A.S., (1984). An analytical model for one-dimensional, three-component condensing
and vaporizing gas drives. Society of Petroleum Engineers Journal, 24(02), 169 –179.
Entov, V.M, (1997). Nonlinear waves in physicochemical hydrodynamics of enhanced oil recovery. Multicomponent
flows. International Conference on Porous Media: Physics, Models, Simulation, Moscow.
Helfferich, F., Klein, G., (1970). Multicomponent Chromatography, Dekker, New York.
Helfferich, F.G., 1981. Theory of multicomponent, multiphase displacement in porous media, SPE Journal, pp. 61–62.
Holden, H., Risebro, N.H. (2016). Front tracking for hyperbolic conservation laws (Vol. 152). Springer.
Isaacson, E.L. 1989. Global solution of a Riemann problem for a non-strictly hyperbolic system of conservation laws
arising in enhanced oil recovery. Enhanced Oil Recovery Institute, University of Wyoming.
Jerauld, G.R., Webb, K.J., Lin, C.Y., Seccombe, J.C. (2008). Modeling low-salinity waterflooding. SPE Reservoir
Evaluation & Engineering, 11(06), 1–000.
Johansen, T., Winther, R., (1989). The Riemann problem for multicomponent polymer flooding. SIAM Journal on
Mathematical Analysis, 20(4), 908 –929.
Johns, R.T., (1992). Analytical theory of multicomponent gas drives with two-phase mass transfer, PhD dissertation,
Department of petroleum engineering, Stanford, California, Stanford University.
SPE-179695-MS 23

Juanes, R., Lie, K.A., (2008). Numerical modeling of multiphase first-contact miscible flows. Part2. Front-tracking/
streamline simulation. Transport in Porous Media, 72, 97(120).
Khorsandi, S., Ahmadi, K., Johns, R.T., (2014). Analytical solutions for gas displacements with bifurcating phase
behavior. SPE Journal, 19(05), 943–955.
Khorsandi, S., Johns, R.T. (2015). Tie-Line Solutions for MMP Calculations By Equations-of-State. In SPE Annual
Technical Conference and Exhibition. Society of Petroleum Engineers.
Khorsandi, S., Shen, W., Johns, R.T., (2015). Global Riemann solver and front tracking approximation of three-component
gas floods, Quarterly of Applied Mathematics, Brown University, American Mathematical Society.
Kipp, K. L., Engesgaard, P., Charlton, S. R. (2004). PHAST--A program for simulating ground-water flow, solute
transport, and multicomponent geochemical reactions. US Department of the Interior, US Geological Survey.
Korrani, A. K. N., Jerauld, G. R., Sepehrnoori, K. (2014). Coupled Geochemical-Based Modeling of Low Salinity
Waterflooding. In SPE Improved Oil Recovery Symposium, Tulsa, Oklahoma (pp. 12–16).
Kozaki, C. (2012), Efficiency of low salinity polymer flooding in sandstone cores. University of Texas at Austin, MSc
dissertation.
Lake, L.W., Johns, R.T., Rossen, W.R., Pope, G.A. (2014). Fundamentals of enhanced oil recovery. Society of Petroleum
Engineers.
Lager, A., Webb, K. J., Black, C. J. J. (2007). Impact of brine chemistry on oil recovery. In 14th European Symposium
on Improved Oil Recovery.
Lager, A., Webb, K., Seccombe, J. (2011). Low salinity waterflood, Endicott, Alaska: Geochemical study & field evidence
of multicomponent ion exchange. In IOR 2011-16th European Symposium on Improved Oil Recovery.
LeVeque, R. J. (2002). Finite volume methods for hyperbolic problems (Vol. 31). Cambridge university press.
Li, W., Dong, Z., Sun, J., Schechter, D. S. (2014, March). Polymer-alternating-gas simulation: A Case Study. In SPE EOR
Conference at Oil and Gas West Asia. Society of Petroleum Engineers.
Luo, H., Al-Shalabi, E. W., Delshad, M., Panthi, K., & Sepehrnoori, K. (2015). A Robust Geochemical Simulator to Model
Improved-Oil-Recovery Methods. SPE Journal.
Mohammadi, H., Jerauld, G. (2012). Mechanistic modeling of the benefit of combining polymer with low salinity water
for enhanced oil recovery. In SPE Improved Oil Recovery Symposium. Society of Petroleum Engineers.
Morrow, N., Buckley, J. (2011). Improved oil recovery by low-salinity waterflooding. Journal of Petroleum Technology,
63(05), 106 –112.
Myint, P. C., Firoozabadi, A. (2015). Thin liquid films in improved oil recovery from low-salinity brine. Current Opinion
in Colloid & Interface Science, 20(2), 105–114.
Orr, F.M., 2007. Theory of gas injection processes. Copenhagen, Tie-Line Publications.
Parkhurst, D. L., Appelo, C. A. J. (1999). User’s guide to PHREEQC (Version 2): A computer program for speciation,
batch-reaction, one-dimensional transport, and inverse geochemical calculations.
PennSim, (2013), Enhanced Oil Recovery Joint Industry Project, Director: Dr. Russell T. Johns, EMS Energy Institute,
The Pennsylvania State University, University Park, PA.
Pires, A.P., Bedrikovetsky, P.G., Shapiro, A.A., (2006). A splitting technique for analytical modelling of two-phase
multicomponent flow in porous media. Journal of Petroleum Science and Engineering, 54(67).
Pope, G.A., 1980. The application of fractional flow theory to enhanced oil recovery. Society of Petroleum Engineers
Journal, 20(03), 191–205.
Pope, G.A., Lake, L.W., Helfferich, F.G. (1978). Cation Exchange in Chemical Flooding: Part 1--Basic Theory Without
Dispersion. Society of Petroleum Engineers Journal, 18(06), 418 –434.
Qiao, 2015, General Purpose Compositional Simulation for Multiphase Reactive Flow with a Fast Linear Solver. PhD
dissertation, Pennsylvania State University.
Qiao, C., Li, L., Johns, R.T., Xu, J. (2015a). A Mechanistic Model for Wettability Alteration by Chemically Tuned
Waterflooding in Carbonate Reservoirs. SPE Journal.
Qiao, C., Li, L., Johns, R.T., Xu, J. (2015b). Compositional Modeling of Dissolution-Induced Injectivity Alteration During
CO 2 Flooding in Carbonate Reservoirs. SPE Journal.
Qiao, C., Johns, R.T., Li, L. (2016). Modeling Low Salinity Waterflooding in Chalk and Limestone Reservoirs. Energy
& Fuels.
RezaeiDoust, A., Puntervold, T., Austad, T. (2011). Chemical verification of the EOR mechanism by using low
saline/smart water in sandstone. Energy & Fuels, 25(5), 2151–2162.
Rossen, W.R., Venkatraman, A., Johns, R.T., Kibodeaux, K.R., Lai, H., Tehrani, N.M. (2011). Fractional flow theory
applicable to non-Newtonian behavior in EOR processes. Transport in porous media, 89(2), 213–236.
24 SPE-179695-MS

Seccombe, J.C., Lager, A., Webb, K.J., Jerauld, G., Fueg, E. (2008). Improving Wateflood Recovery: LoSalTM EOR Field
Evaluation. In SPE Symposium on Improved Oil Recovery. Society of Petroleum Engineers
Shaker Shiran, B., Skauge, A. (2013). Enhanced oil recovery (EOR) by combined low salinity water/polymer flooding.
Energy & Fuels, 27(3), 1223–1235.
Sheng, J.J. (2010). Modern chemical enhanced oil recovery: theory and practice. Gulf Professional Publishing.
Sheng, J.J. (2014a). A comprehensive review of alkaline–surfactant–polymer (ASP) flooding. Asia-Pacific Journal of
Chemical Engineering, 9(4), 471–489.
Sheng, J. J. (2014b). Critical review of low-salinity waterflooding. Journal of Petroleum Science and Engineering, 120,
216 –224.
Sheng, J. J., Leonhardt, B., Azri, N. (2015). Status of Polymer-Flooding Technology. Journal of Canadian Petroleum
Technology, 54(02), 116 –126.
Sorbie, K. S. (2013). Polymer-improved oil recovery. Springer Science & Business Media.
Valocchi, A. J., Street, R. L., & Roberts, P. V. (1981). Transport of ion-exchanging solutes in groundwater: Chromato-
graphic theory and field simulation. Water Resources Research, 17(5), 1517–1527.
Venkatraman, A., Hesse, M.A., Lake, L.W., Johns, R.T. (2014). Analytical solutions for flow in porous media with
multicomponent cation exchange reactions. Water Resources Research, 50(7), 5831–5847.
Vermolen, E., Van Haasterecht, M.J., Masalmeh, S.K., Faber, M.J., Boersma, D.M., Gruenenfelder, M.A. (2011). Pushing
the envelope for polymer flooding towards high-temperature and high-salinity reservoirs with polyacrylamide based
ter-polymers. In SPE Middle East Oil and Gas Show and Conference. Society of Petroleum Engineers.
Vledder, P., Gonzalez, I. E., Carrera Fonseca, J. C., Wells, T., Ligthelm, D. J. (2010). Low salinity water flooding: proof
of wettability alteration on a field wide scale. In SPE Improved Oil Recovery Symposium. Society of Petroleum
Engineers.
Voegelin, A., Vulava, V. M., Kuhnen, F., Kretzschmar, R. (2000). Multicomponent transport of major cations predicted
from binary adsorption experiments. Journal of contaminant hydrology, 46(3), 319 –338.
Wagner, D.H. (1987). Equivalence of the Euler and Lagrangian equations of gas dynamics for weak solutions. Journal of
differential equations, 68(1), 118 –136.
Walsh, M.P., Lake, L.W. (1989). Applying fractional flow theory to solvent flooding and chase fluids. Journal of
Petroleum Science and Engineering, 2(4), 281–303.

Potrebbero piacerti anche