You are on page 1of 17

J170966 DOI: 10.

2118/170966-PA Date: 6-August-15 Stage: Page: 767 Total Pages: 17

A Mechanistic Model for Wettability


Alteration by Chemically Tuned
Waterflooding in Carbonate Reservoirs
C. Qiao, L. Li, R.T. Johns, and J. Xu, Pennsylvania State University

Summary coreflooding experiments (Webb et al. 2004; McGuire et al. 2005;


Injection of chemically tuned brines into carbonate reservoirs Lager et al. 2006). Incremental-oil recovery by up to 40% OOIP
has been reported to enhance oil recovery by 5–30% original oil has been demonstrated in carbonate cores (Zhang et al. 2007;
in place (OOIP) in coreflooding experiments and field tests. One Yousef et al. 2010). Incremental-oil recoveries from field tests,
proposed mechanism for this improved oil recovery (IOR) is however, are generally smaller than those from corefloods.
wettability alteration of rock from oil-wet or mixed-wet to more- Increases of 15% OOIP have been reported in sandstone reser-
water-wet conditions. Modeling of wettability-alteration experi- voirs (Webb et al. 2004). Oil recovery of 50% OOIP by use of
ments, however, is challenging because of the complex interac- seawater injection in carbonate reservoirs, such as in the Ekofisk
tions among ions in the brine and crude oil on the solid surface. field in the North Sea reservoir, have been reported (Hallenbeck
In this research, we developed a multiphase and multicomponent et al. 1991; Austad et al. 2008; Yousef et al. 2012c).
reactive transport model that explicitly takes into account wett- Wettability alteration of the rock from oil-wet to water-wet
ability alteration from these geochemical interactions in carbon- has been suggested as the primary mechanism for increased oil re-
ate reservoirs. covery during low-salinity waterflooding in carbonates (Morrow
Published experimental data suggest that desorption of acidic- 1990; Buckley and Liu 1998; Austad et al. 2012). Oil recovery is
oil components from rock surfaces make carbonate rocks more generally greater in water-wet reservoirs because of the higher oil
water-wet. One widely accepted mechanism is that sulfate (SO2 4 )
mobility owing to its lower affinity to rock surfaces. Water break-
replaces the adsorbed carboxylic group from the rock surface, through is typically slower in water-wet rocks compared with oil-
whereas cations (Ca2þ, Mg2þ) decrease the oil-surface potential. wet reservoirs. In addition, in fractured rocks, a water-wet matrix
In the proposed mechanistic model, we used a reaction network allows for water imbibition and countercurrent flow of oil. Most
that captures the competitive surface reactions among carboxylic carbonate reservoirs are not completely oil-wet; instead, the rocks
groups, cations, and sulfate. These reactions control the wetting usually have mixed wettability depending on the nature of the
fractions and contact angles, which subsequently determine the cap- mineral surface, oil properties, and fluid/rock interactions (Ander-
illary pressure, relative permeabilities, and residual oil saturations. son 1987; Morrow 1990; Peters 2012). The relative proportion of
The developed model was first tuned with experimental data oil-wet and water-wet surfaces determines the overall capillary
from the Stevns Klint chalk and then used to predict oil recovery pressure, relative permeability, and residual oil saturation (ROS),
for additional untuned experiments under a variety of conditions which ultimately control oil recovery (Anderson 1987; Ustohal
where IOR increased by as much as 30% OOIP, depending on sa- et al. 1998; Delshad et al. 2003; O’Carroll et al. 2005). Low-salin-
linity and oil acidity. The numerical results showed that an ity seawater has been found to increase the proportion of the
increase in sulfate concentration can lead to an IOR of more than water-wet surface during spontaneous imbibition experiments in
40% OOIP, whereas cations such as Ca2þ have a relatively minor the Stevns chalk (Strand et al. 2006a, 2006b; Puntervold and Aus-
effect on recovery (approximately 5% OOIP). Physical parame- tad 2008; Strand et al. 2008; Puntervold et al. 2009). Favorable
ters, including the total surface area of the rock and the diffusion contact-angle hysteresis was observed during injection of low-sa-
coefficients, control the rate of recovery, but not the final oil re- linity brine containing sulfate (Alotaibi et al. 2010; Yousef et al.
covery. The simulation results further demonstrate that the opti- 2010; Gupta and Mohanty 2011; Yousef et al. 2011; Yousef et al.
mum brine formulations for chalk are those with relatively 2012a, 2012b, 2012c). Incremental-oil recovery differed signifi-
abundant SO2 4 (0.096 mol/kg water), moderate concentrations of cantly from 0 to 40% OOIP under various experimental condi-
cations, and low salinity (total ionic strength of less than 0.2 mol/ tions (Fathi et al. 2010).
kg water). These findings are consistent with the experimental Possible mechanisms for the observed wettability alteration
data reported in the literature. The new model provides a powerful include fine-particle migration (Tang and Morrow 1999), ion
tool to predict the IOR potential of chemically tuned waterflood- exchange (Lager et al. 2006), mineral dissolution (Hiorth et al.
ing in carbonate reservoirs under different scenarios. To the best 2010), and sorption and desorption of carboxylic groups (Zhang
of our knowledge, this is the first model that explicitly and mecha- et al. 2007). For Stevns chalk, spontaneous imbibition and chro-
nistically couples multiphase flow and multicomponent surface matographic wettability tests verified that SO2 2þ 2þ
4 , Ca , and Mg
complexation with wettability alteration and oil recovery for car- ions actively participate in surface reactions that alter wettability
bonate rocks specifically. (Strand et al. 2003, 2006a; Zhang et al. 2007). Austad et al.
(2008) suggest that sulfate adsorption on positively charged chalk
Introduction surfaces and desorption of the carboxylic group from the surface
Changing the ionic composition of injection water during water- reduces the affinity of the surface to oil (RezaeiDoust et al. 2009).
flooding has been reported to lead to improved oil recovery (IOR) According to this mechanism, experimental studies reported the
in recent years (Yildiz and Morrow 1996; Lager et al. 2006; You- optimal ionic composition for IOR in carbonates (Fathi et al.
sef et al. 2010). Increases in oil recovery between 5 and 38% orig- 2011). Other factors, including temperature, oil composition, and
inal oil in place (OOIP) have been observed in sandstone- water-phase composition, are also observed to play an important
role in determining oil recovery (Hjelmeland and Larrondo 1986;
Strand et al. 2006b; Zhang et al. 2007; Puntervold and Austad
Copyright V
C 2015 Society of Petroleum Engineers
2008; Strand et al. 2008; Puntervold et al. 2009; Fathi et al. 2011).
This paper (SPE 170966) was accepted for presentation at the SPE Annual Technical Significant advances have been made in recent years to predict
Conference and Exhibition, Amsterdam, 27–29 October 2014, and revised for publication.
Original manuscript received for review 4 August 2014. Revised manuscript received for
wettability alteration and oil recovery (Høgnesen et al. 2006; Jer-
review 25 November 2014. Paper peer approved 11 February 2015. auld et al. 2008; Yu et al. 2009; Evje and Hiorth 2011; Andersen

August 2015 SPE Journal 767

ID: jaganm Time: 19:17 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 768 Total Pages: 17

Multiphase Flow This paper is organized as follows. We first introduce the gen-
eral multiphase-flow and reactive-transport equations, and then
+ · = 0
present the reaction network and the relationships between
chemical concentrations, capillary pressure, and relative perme-
+ · = 0
ability. We then focus on the model validation with the base-
case experimental data and sensitivity analysis of the important
Wettability, relative parameters and processes in determining wettability and the oil-
Solute transport permeability, and capillary recovery factor.
,→ pressure alteration
, ,
Methodology
Reactive Transport In this section, we introduce the general multiphase-flow and re-
active-transport equations, the reaction network, and the wettabil-
+ + · + =0 ity-alteration model. The finite-difference-solution approach is
presented at the end of this section.
Low-salinity flooding involves both multiphase flow and geo-
chemical reactions. The injected brine has ionic compositions dif-
Fig. 1—Schematic of the governing equations for multiphase ferent from the formation water. This difference perturbs the
flow (oil and water) and multicomponent reactive transport. The
original thermodynamic equilibrium and leads to surface-geo-
two processes interact by means of solute transport and wett-
ability alteration. chemical reactions, which alters the concentrations of surface spe-
cies and potentially the wettability. The wettability controls
capillary pressure and relative permeability, which in turn affect
et al. 2012). Jerauld et al. (2008) proposed a fully compositional multiphase flow and recovery.
model that included the transport of salts in the aqueous phase as We have developed an implicit-pressure/explicit-concentration-
an additional single-lumped component. They determined the formulated code (PennSim Toolkit 2013) to solve the coupled mul-
relationship between the relative permeability and ROS, but from tiphase-transport and chemical-reaction equations in this research
linear interpolation of the wetting state by use of the salinity with- (Fig. 1). The mass-conservation equations for oil and water phases
out tracking individual species. Yu et al. (2009) and Andersen are solved for the pressure and saturation sequentially. The water-
et al. (2012) assumed a single wetting agent that modified the saturation and water-phase flux from the solutions of the multi-
rock wettability through adsorption. The simplification of wett- phase-flow equations are used in the reactive-transport equations.
ability alteration by including one or two chemical species is not Reactive-transport equations are then solved sequentially for the
sufficient to capture the complex interactions among multiple spatial/temporal evolution of the concentrations of aqueous and
components in water, oil, and solid surfaces. Brady et al. (2012) surface species. The multiphase flow and reactive transport are
used a surface complexation model with reaction networks rele- linked through the interactions among surface reactions, wettabil-
vant to carbonate rocks and sandstones (Brady and Krumhansl ity, relatively permeability, and capillary pressure.
2012; Brady et al. 2012). However, their reaction approach has
not been coupled with multiphase flow to understand dynamic Multiphase-Flow Equations. The mass-conservation equations
effects on wettability alteration. Other models with multiple of the immiscible oil- and water-fluid phases are as follows:
chemical reactions either assumed that mineral dissolution modi-
fies wettability (Evje and Hiorth 2011; Andersen et al. 2012), or @
were designed only for sandstones where cation exchange was ð/Sa qa Þ þ r  ðqa~
u a Þ ¼ 0; a ¼ o; w; . . . . . . . . . . ð1Þ
@t
believed to be the mechanism of wettability alteration (Dang et al.
2013). In general, there is currently a lack of a detailed representa- where / is porosity (dimensionless) and S, q, and ~u are the satura-
tion of the surface-geochemical reactions and the corresponding tion (dimensionless), fluid density (kg/m3), and volumetric flow
wettability alterations in multiphase-flow models. No existing rates (m3/s), respectively, for the oil and water phases. The sub-
models can be used to quantitatively link water and surface reac- script w is for the water phase, and the subscript o is for the oil
tions, wettability, capillary pressure, relative permeability, and phase. Darcy’s law governs the flow rate of different phases:
(eventually) oil recovery, or to predict oil recovery under different
injection-water compositions and conditions for carbonate rocks. kkra
ua ¼
~ rðPa  qa gZÞ; . . . . . . . . . . . . . . . . . . . . . ð2Þ
In the geochemistry community, multicomponent reactive- la
transport models have been developed since the 1980s (Lichtner
1985; Steefel et al. 2005) and have been extensively used to where k is absolute permeability (m2); Z is the depth (m); and l,
understand and predict subsurface reactive-transport processes in g, and P are the viscosity (cp), gravitational constant (m/s2), and
many applications (Davis et al. 2004; Li et al. 2010, 2011). Appli- the pressure of the fluid phase (Pa), respectively. The pressure dif-
ference between oil and water phases is the capillary pressure:
cations of these complex surface reactions in IOR processes,
although promising from a geochemistry point of view, have not
Pcow ¼ Po  Pw : . . . . . . . . . . . . . . . . . . . . . . . . . . . ð3Þ
yet been made to the best of our knowledge.
In this research, we develop a model that couples multiphase
The capillary pressure Pcow and the relative permeabilities krw
flow with detailed, mechanistic understanding of surface reactions
and kro depend on water saturation, pore structure, and rock wett-
to systematically investigate the complex interactions among mul-
ability. The saturation relation completes the set of equations:
ticomponent surface reactions, wettability, and oil recovery. We
propose a reaction network for carbonates based on a double sur-
So þ Sw ¼ 1: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð4Þ
face complexation model (Brady et al. 2012a, 2012b). The new
model was tuned with data from a low-salinity seawater-imbibi-
The unknowns for the multiphase-flow system are the pressure
tion experiment, where both porous-media properties (porosity,
and saturation of the different fluid phases.
permeability, capillary pressure, and relative permeability) and
geochemical reactions (aqueous- and surface-complexation reac-
tions) play important roles. Simulations were carried out with this Reactive-Transport Equations. Reactive-transport equations
new model under an array of conditions to understand the control- describe the coupled process of solute transport and reactions.
ling parameters during the chemically tuned waterfloods. By Compared with standard reactive-transport models for water-satu-
chemically tuned waterflooding, we refer to the injection of water rated porous media (Steefel and Lasaga 1994), PennSim has vary-
that is adjusted in the ionic composition. ing water saturation and three interfaces (oil/water, oil/solid, and

768 August 2015 SPE Journal

ID: jaganm Time: 19:17 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 769 Total Pages: 17

(a) (c)

O
w

(b) Oil Water

Water

Solid Solid

Carboxylic group or other surface active components


Surfate
Divalent Cation like Ca2+ and Mg2+

Fig. 2—Illustration of the multiphase reaction network in a pore of mixed wettability. (a) A mixed-wet pore by use of a triangle
model; o 5 oil; w 5 water. (b) The ion-species distribution on a water-wet surface. (c) The chemical-ion species at the interfaces
between crude oil, water, and solid.

water/solid interfaces). The species are partitioned into primary and has been shown to be crucial in low-salinity flooding (Zhang
and secondary species. The partition is determined in such a way and Austad 2005a). Here we use AN to quantify the amount of
that the concentrations of the secondary species can be explicitly carboxylic acids in the oleic phase.
expressed by those of the primary species through the mass-action The carboxylic acids represent the polar oil components. As
law (Lichtner et al. 1996). The mass-conservation equation for the shown in Reactions 1 through 3 in Table 1, the carboxylic acids
primary species p is dissociate at the oil/water interface and react with ions in the water
! ! phase. The species at the oil/water interface occupies an oil-sur-
@ XNsec X
Nsec
face site, following Brady et al. (2012). A diffusive-layer model is
Mp þ qp Mq þ r  Fp þ qp Fq ¼ 0; used to quantify the activity of the surface species and the electro-
@t q¼1 q¼1
static forces (Dzombak and Morel 1990). The expression for the
p ¼ 1; …; Np ;                       ð5Þ equilibrium constant of Reaction 2 is shown as an example:
 
where subscripts p and q represent the primary species p and sec- Fwo
exp ½COOCaþ aHþ
ondary species q and qp represents the ðq; pÞ entry of the stoichi- RT
ometry coefficient matrix. The number of primary species Np equals K2;eq ¼ ;
½COOHaCa2þ
Ntot  Nsec , where Ntot is the total number of species and Nsec is the
number of secondary species. The definition of molar density M, where ½COOH and ½COOCaþ  are the surface concentrations
flux F, and the derivation of Eq. 5 are in Appendix A. The set of (mol/m2) of carboxylic acid and the surface complex of carbox-
unknowns for the reactive-transport equations includes the aqueous- ylic calcium, respectively; aHþ is the activity of Hþ in aqueous
species concentration Ciw , the solid-surface-species concentration phase (dimensionless); F is Faraday’s constant (9.648 104 C/
Cis , and the oil-surface-species concentration Cio , where the sub- mol); wo is the oil/water-interface charge potential (mV); R is the
script i represents a species. Details of the reactive-transport-model- gas constant (8.314 J/Kmol); and T is the absolute temperature
ing formulation can be found in Yeh and Tripathi (1991), Steefel (Kelvin). The surface potential is calculated by use of the Gouy-
and Lasaga (1994), and Walter et al. (1994). The system of general Chapman theory that relates the surface-charge density to surface
equations is coupled with the mass-action law discussed later. potential in the following form (Gouy 1910; Chapman 1913):
 
p Fw
Multiphase Reaction Network. As illustrated in Fig. 2, the ro ¼  8e0 em IRTsinh  o ; . . . . . . . . . . . . . . . ð6Þ
2RT
reaction network includes the interactions among Ca2þ, Mg2þ,
SO2
4 , and absorbed-oil species. The reaction network includes where ro is the charge 2
aqueous reactions and surface-complexation reactions at the oil/ Xdensity at the oil/water interface (C/m )
water, solid/oil, and solid/water interfaces (Brady and Krumhansl calculated from ro ¼ z C , where zi is the charge carried by
i i i;o
2012; Brady et al. 2012). The reactions on the solid/water inter- ion species i; e0 is the dielectric constant of water (55.3,
face include the adsorption of sulfate and the carboxylic group. dimensionless); em is the permittivity of free space (8.8541013
The reactions on the oil/water interface include the dissociation C V 1 dm1 ); and I is the ionic strength of water (mol/kg water).
of carboxylic acids and reactions between the carboxylic group The solid-surface potential can be calculated in a similar manner.
and multivalent cations. All aqueous reactions are fast reactions The calculation follows the same method of Hiorth et al. (2010).
and are assumed to be at equilibrium and are controlled by reac- This surface-complexation model integrates the effects of surface
tion thermodynamics (Lichtner et al. 1996; Langmuir 1997). charge, solution ionic strength, temperature, and surface potential.
The concentrations of the polar components in crude oil The reactions between brine and the calcite surface are repre-
largely affect the initial wettability and the wettability alteration sented by Reactions 4 and 5 in Table 1. The species > CaOH rep-
by seawater (Strand et al. 2003; Fathi et al. 2011). For refined oil resents the reactive site on the calcite surface because > CaOHþ 2
where polar components are removed, no low-salinity-water or was found to sorb strongly on the oil surface (Brady et al. 2012).
seawater improved-oil-recovery (IOR) effect has been observed Reaction 4 describes the hydration of the calcite-surface site. The
(Hirasaki and Zhang 2004; Robertson 2007). The acid number equilibrium constant is known to highly depend on temperature
(AN) quantifies the abundance of polar components in crude oil (Austad et al. 2008; Fathi et al. 2010). Here we use the data

August 2015 SPE Journal 769

ID: jaganm Time: 19:17 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 770 Total Pages: 17

where the activity coefficients were calculated according to the


extended Debye-Hückel model (Helgeson et al. 1970):
pffiffi
Az2i I
lnci ¼  pffiffi þ bI;
1 þ a0i B I

where the parameters A, B, and b are temperature-dependent pa-


rameters taken from the EQ3 database (Wolery et al. 1990). The
parameter a0i is the ion size, and I is the ionic strength of the brine.
Note that with higher ionic strength, the activity coefficients are
smaller, which leads to lower activity of specific species.

Wettability-Alteration Model. Most carbonate reservoirs are


classified to have mixed wettability (Buckley et al. 1996; Helland
and Skjaeveland 2006). Here, the rock surface is considered as
containing both water-wet and oil-wet surfaces. Different surfaces
can have different contact angles. Both receding-contact angles
(altered, water-wet) and advancing-contact angles (unaltered, oil-
wet) varied in core experiments (Buckley et al. 1996; Drummond
and Israelachvili 2004; Zhang and Austad 2005a; Alotaibi et al.
2010). According to Young’s equation, the contact angle has the
Table 1—The reaction network and equilibrium constants. Here, “–”
following relation derived from the force balance:
means species at the oil/water interface and “>” means species on
the solid surface. All other species are in the aqueous phase. All cos  cws
reactions are assumed at equilibrium. cosh ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . ð7Þ
cow

where cws , cos , and cow are the interfacial tension (IFT) (mN/m)
interpolated from Evje and Hiorth (2011). Similar to the oil/ between water and solid, between oil and solid, and between oil
water-interface reactions, the equilibrium constants of Reaction 5 and water, respectively. The contact angle describes how much a
can be written as mineral surface prefers one phase to another and is a result of the
  three-phase (brine/oil/surface) interaction. The IFTs are deter-
Fw mined by the surface concentration of ion species through Gibbs
exp  s ½> CaSO 4
RT equation (Gibbs 1948). The contact angle is then a function of sur-
K5 ¼ ; face concentrations on the interface. We model the contact angle
½> CaOHþ 2 aSO2
4 as a linear function of surface concentration for the oil-wet and
water-wet surfaces of carbonates, which is justified in Appendix
where ws is the solid/water-interface potential (mV), which can B, where the derivation is modeled after the Gibbs isotherm. The
be calculated similarly to wo . Experimental results show that linear interpolation is shown in Eqs. 8 and 9:
Ca2þ, Mg2þ, and SO2 4 control surface potential, but Ca2þ or
Mg2þ alone without SO2 4 cannot alter wettability (Zhang et al. ½> CaSO 4
2007). SO2 alone, however, cannot alter wettability either coshw ¼ coshw;0 þ ðcoshw; 1  coshw; 0 Þ;
4 ½> Ca; total
(Strand et al. 2006a; Zhang et al. 2007). Reactions 1 through 5
                   ð8Þ
show that SO2 4 determines the potential at the solid surface,
whereas Mg2þ and Ca2þ complex with desorbed carboxylic acids ½> CaOHþ 
2 ðCOO Þ
from the solid/fluid interface, which allows sufficient change to cosho ¼ cosho;0 þ ðcosho;1  cosho;0 Þ;
½> Ca; total
occur. These reactions describe the different roles of aqueous spe-
cies. Reaction 6 in Table 1 represents the acid/base interactions                    ð9Þ
between oil/water interface and solid surface. Ion binding is not
included here because on calcite surfaces a positive surface where ½> Ca; total represents the total concentration of the sur-
charge dominates below a pH of 9.4 (Buckley and Liu 1998), face site on the carbonate surface. The contact angle hw on the
whereas the experiment in Fathi et al. (2010) used a pH value of water-wet surface (receding-contact angle) is interpolated by the
8.4. The equilibrium constant for Reaction 6 was obtained by his- concentration of the water-wet agent (sulfate). The oil-wet contact
tory matching of the base-case scenario. angle ho (advancing-contact angle) is interpolated by the oil-wet
Reactions 1 through 6 account for the electrostatic interactions agent (carboxylic group). The contact angles hw; 0 ; hw; 1 ; ho ;0 , and
between the oil/water interface and the solid surface. The crude- ho;1 are input values representing the extreme cases (0 and 90 for
oil/brine interface is known to carry a negative charge, as repre- water-wet, 90 and 150 for oil-wet).
sented in Reaction 1, whereas a chalk surface is known to be posi- The surface concentration also determines the oil-wet and
tively charged, as shown in Reaction 4 (Hiorth et al. 2010). The water-wet surface fractions. The water-wet fraction (WWF) is cal-
electrostatic attraction between opposite charges leads to the oil culated as a linear function of the surface-site concentration as
sorption on the solid surface (Nasralla and Nasr-El-Din 2012).
½> CaOHþ2 ðCOO Þ

Those reactions represent the competitive adsorption of carbox- WWF ¼ 1  :
ylic acids and SO2 ½> Ca; total
4 on the chalk/water interface and the competi-
tive compounding of Ca2þ, Mg2þ, and >CaOH 2 on the oil/water The change in relative permeability and capillary pressure
interface. Reactions 7 through 9 are aqueous-speciation reactions
functions have typically been determined by linear interpolation
that are also important in determining pH.
with respect to a wettability index (Delshad et al. 2009; Yu et al.
For all aqueous species (Ca2þ, Mg2þ, Hþ, and others), the ac-
2009). Our model uses an experimentally verified wettability
tivity was calculated from
index as well as theoretical support from equations by Gibbs
(1948) and Cassie (1948). The capillary pressure function fol-
ai ¼ ci Ci ;
lows the Leveret-Cassie equation (O’Carroll et al. 2005):

770 August 2015 SPE Journal

ID: jaganm Time: 19:17 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 771 Total Pages: 17

10 1
krowater-wet

0.8
5 kroWWF = 0.5

Relative Permeability
Capillary Pressure water-wet

WWF = 0.52 0.6


0 krooil-wet oil-wet
krw
WWF = 0.82 WWF = 0.5
0.4 krw

–5
0.2 water-wet
krw
oil-wet

–10 0
0 0.5 1
0 0.2 0.4 0.6 0.8
(a) Water Saturation (b) Water Saturation

Fig. 3—Illustration of wettability-dependent capillary pressure and relative permeability. (a) Relative permeability for oil-wet
(WWF 5 0), water-wet (WWF 5 1), and intermediate-wettability (WWF 5 0.5) state. (b) Imbibition capillary pressure for oil-wet, water-
wet, and two intermediate states with different WWF values: 0.52 and 0.82. Note that the equilibrium water saturation under capil-
lary pressure is the x-axis intercept. As the rock becomes more water-wet, larger water saturation can be achieved by spontaneous
imbibition.

Pc ðSw ; WWFÞ ¼ WWFcoshw Pww permeability between completely water-wet and oil-wet cases, as
c ðSw Þ
follows:
þ ð1  WWFÞcosho Pow
c ðSw Þ;        ð10Þ
  
krw ¼ WWF krw;ww þ ð1  WWFÞkrw;ow ;
where and Pww
c ðSw Þ Pow
c ðSw Þ
are the capillary pressure functions
for completely water-wet and oil-wet surfaces. Eq. 10 has been 
kro 
¼ WWF kro;ww 
þ ð1  WWFÞkro;ow ;
experimentally verified to provide excellent predictions of the
capillary pressure as a function of saturation (Ustohal et al. 1998; 
where kro;ww 
, krw;ww 
, kro;ww 
, and kwo;ow are the oil and water end-
O’Carroll et al. 2005). Surface roughness has been included in the point relative permeability for the oil-wet or water-wet case. One
completely oil-wet and completely water-wet capillary pressures. set of relative permeability curves is shown in Fig. 3b, with the
One example of the capillary pressure for mixed wettability is corresponding parameters (endpoint relative permeability and
shown in Fig. 3a, where the oil-wet and water-wet data are from Corey exponent) in Table 2. In Fig. 3b, the oil-wet relative per-
Webb et al. (2005). The curves for intermediate wetting state meability function was adapted from Høgnesen et al. 2006.
were obtained by use of Eq. 10. In Fig. 3b, the x-intercept repre- Increasing water wettability shifts the curves to the right, which
sents the maximum water saturation that can be achieved by spon- favors oil-phase flow. The treatment is consistent with experimen-
taneous imbibition. Our model captures the fact that the residual tal findings (Owens and Archer 1971).
oil saturation (ROS) depends on wettability. The simulation code PennSim was used to solve the multi-
The Brooks-Corey formulation was used here to describe the phase-flow equations. PennSim uses a finite-volume discretization
relative permeability as a function of normalized fluid saturation (Fung et al. 1992) and a noniterative implicit-pressure/explicit-
and water-wet fraction (Anderson 1987; Delshad et al. 2009). The concentration method (Coats 2000) to solve the governing equa-
relative permeability was assumed to depend on the endpoint rela- tions. We sequentially obtained the immiscible multiphase-flow
tive permeability and relative saturation using the following solution and then a chemical reactive-transport solution. The pro-
equations: cedure uses an operator-splitting technique with a strict restriction

krw ¼ krw ðS Þnw ; on timestep size (Zysset et al. 1994). The calculation procedure
for one timestep is shown in Fig. 4.

kro ¼ kro ð1  S Þno
Results and Discussion
where the normalized water saturation S was calculated as This section presents a validation of the model by use of the base-
So  Sor case scenario, sensitivity analyses of key parameters, and the
S ¼ ; effects of various brine compositions on wettability and recovery.
1  Swi  Sor
The ultimate recovery in this paper refers to the recovery factor at
where Swi is the initial water saturation and Sor is the ROS. For the end of 40 days, when the oil-production rate is very low. The
the mixed and fractional wettability, the fluids were assumed to discussion provides a mechanistic and quantitative understanding
flow in parallel in both oil-wet and water-wet pore networks. of processes involved in low-salinity flooding and identifies the
Therefore, WWF was also used to linearly interpolate the relative most-important parameters.

Table 2—Corey’s relative permeability model parameters.

August 2015 SPE Journal 771

ID: jaganm Time: 19:17 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 772 Total Pages: 17

Begin one timestep calculation

Pressure Solver
Calculate Jacobian for implicit-pressure solution
Multigrid-preconditioned iterative linear solver

Transport Solver
Calculate flux between blocks using the new pressure and old transmissibility
Explicitly calculate moles of components and ion species in each block

Reaction Solver
Update mole fraction of aqueous species
Perform chemical-equilibrium calculation

Property Update
Calculate contact angle, water-wet fraction, and change wettability-affected functions

End one timestep

Fig. 4—Illustration of the flow chart for one timestep in the implicit-pressure/explicit-saturation solution.

Base-Case Simulations. Various core experiments have been relation in Eq. 3. The fluid initial water composition was the same
performed to understand the mechanism and identify the optimal as the formation water.
conditions for chemically tuned waterflooding (Austad et al. We performed 2D simulations by use of the radial symmetry
2008; Fathi et al. 2010; Yousef et al. 2010, 2012a, 2012b). We of the core, where the core was discretized into 30  100 grid-
used the data from Fathi et al. (2010) to validate our model and to blocks in the r, z coordinates, as shown in Fig. 5. Fig. 6 compares
obtain key parameters. We selected this data set because their the oil-recovery data and the modeling output. The reaction-equi-
experiments were consistently carried out at 110 C with different librium K6 value was tuned to match the base case. Here, we com-
brine compositions while maintaining all other conditions, which pare three methods of different levels of complexity in
allows easy comparison among different cases. reproducing the oil recovery by use of brines of different compo-
The experiments used homogeneous Stevns Klint chalk cores sition. Method A was modeled after Høgnesen et al. 2006, where
(3.8 cm in diameter and 7.0 cm in length) with porosity of approx- fixed capillary pressure and relative permeability functions were
imately 45% and permeability between 1 and 2 md (1015 m2). used, as is typically the case for multiphase-flow simulations. Oil-
The cores were first cleaned with distilled water. After the drying wet functions were used for FW; water-wet functions were used
process, the cores were saturated with formation brine and then for SW0NaCl; and intermediate-wet functions were used for SW.
flooded with oil to establish the initial water saturation. The oil These functions were chosen on the basis of our understanding of
used was diluted from acid-reservoir-stabilized oil with n-heptane the wettability with respect to different water compositions.
to an equivalent acid number of 1.9 mg-KOH/g. The cores were Method B includes the transport and adsorption reaction of SO2 4
aged at 90 C for 4 weeks to restore to the reservoir condition, as the single reacting solute and calculates the wettability as a
where mixed wettability has been established. The cores were function of the adsorbed sulfate mass (Yu et al. 2009). The capil-
then immersed in synthetic brine at 110 C, after which spontane- lary pressure and relative permeability functions were obtained by
ous imbibition began. The produced oil was collected over time, interpolation by use of the equations kr ¼ WWF krww þ ð1 
after which chromatographic wettability tests were performed to WWFÞkrow and Pc ¼ WWF Pww ow
c þ ð1  WWFÞPc , where WWF is
determine the water-wet fraction. Three different synthetic brines calculated from the solid-surface fraction of sulfate ions (i.e.,
were used, including formation water (FW), seawater (SW), and CSO24 ;s
=Ctotal;s ). The sulfate-adsorption reaction was modeled with
seawater depleted in NaCl (SW0NaCl), as listed in Table 3. The the Langmuir isotherm without the effect of salinity. Method C is
comparison among cases allows identification of the important our mechanistic method with PennSim, as discussed in the Method-
ions during the low-salinity-imbibition process. ology section. The reaction system explicitly included the effects
During the spontaneous imbibition, the ions in the immersing of different ions on the reaction-driven wettability alteration.
brine were transported and diffused into the core. The chemical The comparison of oil-recovery curves shown in Fig. 6 shows
reactions altered the wettability and improved imbibition. Sponta- the necessity of including the geochemistry details. In the FW
neous imbibition causes countercurrent flow, during which the case, with fixed wettability, Method A reproduced the countercur-
volumetric flow rate of oil and water were equal, but in opposite rent flow and the subsequent oil production for the FW case.
directions. Spontaneous imbibition is simulated by solving Eqs. 1 However, it overestimated the oil-production rates for SW and
through 5 with Dirichlet boundary conditions. The boundary pres- SW0NaCl when wettability alteration occurred. Method B cannot
sure for both phases was set to be the backpressure; the boundary distinguish between the SW and SW0NaCl cases, as indicated by
saturation for water was 1.0; and the brine concentration on the their overlapping oil-recovery curves. This is because of the over-
boundary was set to be the same as the imbibing fluid. The initial simplified representation by use of a single component and single
oil saturation determined from the experiments was used and the reaction. PennSim accurately matched the oil-recovery curves for
water-phase pressure was the same as the backpressure, whereas the FW and SW0NaCl cases, and predicted approximately the re-
the oil-phase pressure was calculated from the capillary pressure covery curve for SW (Fig. 6c). Our model reproduced the oil-

772 August 2015 SPE Journal

ID: jaganm Time: 19:17 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 773 Total Pages: 17

Table 3—Synthetic-brine compositions used in experiments (Fathi et al. 2010) and simulations. We use
the same notation for different types of brine as those used in Fathi et al. (2010). FW denotes formation
water; SW denotes seawater; SW0NaCl denotes seawater with no NaCl; SW0NaCl0SO4 denotes
seawater with no NaCl or SO2 4 ; SW0NaCl4SO4 denotes seawater with no NaCl and with the sulfate
concentration adjusted to four times that of seawater; and SW0NaCl4Ca denotes seawater with no
NaCl but with the calcium concentration adjusted to four times that in seawater.

recovery decrease with increasing NaCl concentration for SW and bition process because seawater was depleted with sodium chlo-
SW0NaCl. Moreover, our model reproduced the measured wetting- ride. In this scenario, Naþ did not participate in sorption and
area fraction in all experiments. The final water-wet fractions simu- acted as a tracer, the spatial evolution of which was a result of the
lated for FW (0.52), SW (0.67), and SW0NaCl (0.83) were within transport process only. Compared with Naþ, the SO2 4 propaga-
5% of the experimentally measured values through chromato- tion was slower because the transport was delayed by adsorption.
graphic wettability tests (Strand et al. 2006a; Fathi et al. 2010). The order of the propagation rate (So < WWF < CSO2 4
< CNaþ )
Fig. 7 shows the profile evolution for the base-case scenario reflects the driving force of the system: the solute diffusion per-
with SW0NaCl. The core was originally saturated with oil, turbed the original surface-chemical equilibrium, followed by
whereas the SW0NaCl solution was at the outside boundary of the altered wettability by the sulfate adsorption, and later by water
core. Sulfate diffused into the core from the boundary. Over time, imbibition with a positive capillary pressure and oil production.
zones of high sulfate concentration expanded. Correspondingly, The coherent profiles among different variables underscore the
the water-wet fraction of the core also expanded into the core and tightly coupled multiphase and reactive-transport processes.
oil was produced from the boundary because of increasing sponta- Fig. 8 shows the sulfate, water-wet fraction, and oil-saturation
neous imbibition. Oil saturation decreased from the initial profiles for SW. SW has seawater levels of sodium chloride,
80–90% to a range of 30–50% on Day 20. The change occurred whereas SW0NaCl has no sodium chloride. At any particular
faster in the first 10 days, with the oil saturation reduced to time, the water-wet fraction at the altered locations was larger for
approximately 36% at the boundary and up to 70% at the center. SW0NaCl than for SW, because there was more SO2 4 adsorbed
At later times, the oil was produced more slowly, primarily on the surface in the SW0NaCl case than in the SW case. Accord-
because of the decreasing positive capillary pressure. Naþ con- ing to the extended Debye-Hückel model, the activity coefficients
centration was high at early time because of the high initial water of the aqueous species in SW0NaCl were larger because of its
salinity. The brine inside the core became diluted during the imbi- low total ionic strength. The activity coefficients varied with mix-
ing of the formation water and the imbibing fluid. Before mixing,
the activity coefficient of sulfate was 0.57 in SW0NaCl and 0.38
3.8 cm in SW. With the same concentration in SW0NaCl and SW, the ac-
tivity of sulfate was 1.5 times greater in SW0NaCl than in SW.
Therefore, more SO2 4 was adsorbed in the SW0NaCl case and
the surface became more water-wet. Comparison between Fig. 8c
and Figs. 7m through 7p shows that the oil saturation for
SW0NaCl is larger than that with SW. That is, high salinity leads
to less wettability alteration and less oil-recovery improvement.

Sensitivity Analysis. The multiple processes involve a large


number of parameters. A sensitivity analysis was performed to
understand the importance of different processes and parameters
7.0 cm in the SW0NaCl case, and the most-effective fluid in improving
oil recovery was compared with other types of brines in this
research. As shown in Table 4, the tested parameters include spe-
cific surface area and surface-site density of the chalk, diffusion
coefficient, equilibrium constants, and the crude-oil acid number.
Surface-reaction-equilibrium constants are important in deter-
mining the chemical-equilibrium state and thus the surface wett-
ability. The reported equilibrium constants of Reactions 2, 3, 5,
and 6 in the literature vary by orders of magnitude (Zhang and
Sparks 1990; Hiorth et al. 2010; Brady and Krumhansl 2012).
These reaction equilibrium constants are chosen to vary by one to
two orders of magnitude in our sensitivity analysis. Fig. 9a com-
Fig. 5—Geometry of the model used in the experiments and
simulations. The radial symmetry of the core allows for 2D sim- pares the oil-recovery curves generated by use of different equi-
ulations instead of 3D ones. The computational domain has 30 librium constants. A larger K5 means larger affinity of sulfate to
blocks in the radial r-direction and 100 gridblocks in the vertical the surface so that more sulfate can adsorb onto the solid surface.
z-direction. The gray blocks are boundary blocks with Dirichlet With the same surface-site density, this will lead to a larger por-
boundary conditions. tion of sulfate-occupying surface sites, which therefore alters the

August 2015 SPE Journal 773

ID: jaganm Time: 19:17 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 774 Total Pages: 17

Experimental FW
0.6
FW simulation A
FW simulation B
0.5
FW simulation C

Oil Recovery
Experimental SW
0.4
SW simulation A
0.3 SW simulation B
SW simulation C
0.2 Experimental SW0NaCl
SW0NaCl simulation A
0.1 SW0NaCl simulation B
SW0NaCl simulation C
0
0 10 20 30 40
Time (days)

Fig. 6—Comparison between experimental oil recovery and simulation results for FW, SW, and SW0NaCl by use of three different
simulation methods. The experimental data are the same for all cases and are represented by use of the symbols. Simulation A
used fixed-wettability functions adapted from Høgnesen et al. 2006. Simulation B used a single adsorption reaction adapted from
Yu et al. (2009). Simulation C used the model developed in this paper. Note that with Simulation B, the curve for SW overlaps with
the curves for SW0NaCl. Only (c) matches the trend in the experimental recovery curve for seawater.

rock surface to a more water-wet state. As such, a larger capillary sulfate for the same proportion of the water-wet fraction. To reach
pressure is expected with a faster and greater oil recovery. Fig. 9a the same wettability, more sulfate needs to be adsorbed, which
shows that with a K5 value one order of magnitude larger, the ulti- takes a longer time. TSS, however, only controls the quantity of
mate oil recovery increased by 5% original oil in place (OOIP). In adsorbed mass and has no effect on the reaction equilibrium. The
addition, the recovery rate was faster because the increased water extent of wettability alteration is the same with different TSS val-
wettability increased the capillary pressure to larger positive val- ues, which leads to the same ultimate oil recovery.
ues. When the K5 value is one magnitude smaller than the base The oil contains the carboxylic group (COOH); larger acid
case, wettability alteration was negligible, which leads to zero numbers (ANs) mean more COOH to react with Ca2þ and
incremental-oil recovery. Mg2þ and other aqueous species. The literature reports a range of
Larger equilibrium constants typically correspond to those surface-density values between 3.1/nm2 and 0.3/nm2, correspond-
under high-temperature conditions. Our results here are consistent ing to acid numbers between 2.07 and 0.17 mg-KOH/g. Fig. 11a
with the findings of the experimental results that low-salinity sea- shows greater recovery with the smaller AN. Fig. 11b shows the
water is more effective under high-temperature conditions (Zhang water-wet fraction along the core radius for different ANs in Fig.
et al. 2006; Strand et al. 2008; RezaeiDoust et al. 2009). 11a. For the small AN, the core was altered to almost completely
The strength of the oil/rock bond was represented by K6 water-wet (water-wet fraction equals 1) at the end of 20 days. For
because Reaction 6 represents the interaction between the carbox- the large-AN case, the water-wet fraction was approximately
ylic group and chalk surface. A large K6 value means a larger 0.82. This difference leads to increases in oil recovery. Such dif-
tendency for oil to attach onto the solid. Fig. 9b shows that ferences can be explained by use of Reactions 1 and 6. Reaction 1
smaller K6 values lead to much-greater oil recovery. The close- shows that larger COOH concentration leads to larger Hþ and
ness between curves for the small and medium cases indicates –COO concentrations, which results in an oil surface that is
that when the K6 is small enough (less than 2.5104), the recov- more negatively charged. With Reaction 6, the increased –COO
ery is not sensitive to the K6 value. concentration leads to more > CaOH2 (COO), which results in a
Figs. 9c and 9d compare the oil recovery with different K2 and more oil-wet chalk surface and less oil recovery. Therefore, a
K3 values. The cations Ca2þ and Mg2þ compound with the car- smaller AN leads to a higher oil recovery (68% OOIP) than a
boxylic group on the oil surface and make the oil-surface charge larger AN (59% OOIP). This is consistent with experimental
more neutral. The decrease of the negative charge enhances the observations in the literature (Zhang and Austad 2005a; Strand et
oil release from the positively charged rock surface. Larger K2 al. 2006a). If the oil does not contain acid at all, the rock is com-
and K3 values result in a stronger tendency of Ca2þ and Mg2þ to pletely water-wet and the spontaneous imbibition has the best re-
attach onto the oil surface. Figs. 9c and 9d show negligible effects covery. However, in this case the salinity of water does not affect
of K2 and K3. Ca2þand Mg2þ have similar roles in the reaction the oil recovery. To our knowledge, the dependence on oil acidity
system. Because there are sufficient divalent cations in the sea- was not included in any other low-salinity models.
water, the cation-compounding reaction is unlikely to be a limit- The impact of diffusion coefficient is shown in Fig. 12a. All
ing reaction. Therefore, Reactions 2 and 3 play a relatively minor diffusion coefficients lead to the same oil recovery on Day 80.
role compared with Reactions 5 and 6. The diffusion coefficient is important because diffusion controls
Fig. 10 shows the effect of the total number of surface sites the transport rate of the ions into the rock. Fig. 12b shows the cor-
(TSS) on oil recovery. TSS was calculated as the product of responding sulfate-concentration profile on Day 10. Sulfate was
chalk-specific-surface area (SSA, in units of m2/g) and surface transported faster with a larger diffusion coefficient, which led to
site density (SSD, which represents the number of sites/m2). Fig. more adsorbed sulfate earlier, and faster wettability alteration and
10 was obtained by varying the SSA with the same SSD. Varying oil recovery. However, the diffusion coefficient does not affect
SSD with the same SSA produces exactly the same curve. As the ultimate wettability and oil recovery at 40 days.
observed from Fig. 10, although the ultimate oil recovery is the From the sensitivity analysis, the parameters are classified into
same for all cases, a smaller TSS leads to faster oil recovery Type I and Type II according to their role in controlling the oil-
whereas a larger TSS leads to slower oil recovery. This is because recovery curve. Type I parameters control the ultimate amount of
of the difference in sulfate adsorption and transport. A larger TSS oil recovery. These include the chemical-equilibrium constants
value means a larger number of surface sites to be occupied by and crude-oil AN. Type II parameters, including the solid-

774 August 2015 SPE Journal

ID: jaganm Time: 19:17 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 775 Total Pages: 17

SO42– Solid-Surface Concentration (μmol/m2)


(a) 0.5 days (b) 5 days 10 days (c) 20 days (d)
mm mm mm mm
2.5
10 10 10 10
20 20 20 20 2
30 30 30 30 1.5
40 40 40 40
1
50 50 50 50
60 60 60 60 0.5

70 70 70 70 0
0 10 mm 0 10 mm 0 10 mm 0 10 mm

Na+ Aqueous Concentration (mol/kg water)

(e) 0.5 days (f) 5 days 10 days (g) 20 days (h) 1


mm mm mm mm
10 10 10 10 0.8
20 20 20 20
30 30 30 30 0.6
40 40 40 40
50 50 50 50 0.4

60 60 60 60
0.2
70 70 70 70
0 10 mm 0 10 mm 0 10 mm 0 10 mm

Water-Wet Fraction
(i) 0.5 days (j) 5 days 10 days (k) 20 days (l)
mm mm mm mm 0.8
10 10 10 10
0.75
20 20 20 20
0.7
30 30 30 30
40 40 40 40 0.65

50 50 50 50 0.6
60 60 60 60
0.55
70 70 70 70
0 10 mm 0 10 mm 0 10 mm 0 10 mm

Oil Saturation
0.9
(m) 0.5 days (n) 5 days 10 days (o) 20 days (p)
mm mm mm mm
0.8
10 10 10 10
20 20 20 20 0.7

30 30 30 30 0.6
40 40 40 40
50 50 50 50 0.5

60 60 60 60 0.4
70 70 70 70
0 10 mm 0 10 mm 0 10 mm 0 10 mm

Fig. 7—Spatial and temporal evolution of the SO22 1


4 , Na concentrations, water-wet fraction, and oil saturation for the imbibition of
SW0NaCl. The horizontal axis is the radial distance from the center of the cylinder core, and the vertical axis is the vertical distance
from the base.

surface-site density and diffusion coefficient, control the rate of with different SO2
4 and Ca concentrations for the same core
oil recovery. Most of these controlling variables (equilibrium con- and crude oil. Fig. 13 shows oil-recovery curves for all cases.
stants, oil acidity, and the diffusion coefficient) are temperature- Among these five cases, the SW0NaCl4SO4 case has the greatest
dependent, indicating the importance of temperature in the effec- oil recovery of approximately 65% OOIP.
tiveness of the chemically tuned waterfloods in carbonates. The SW0NaCl0SO4 case behaves the same as formation
water, indicating no improved oil recovery without the pres-
Effect of Brine Composition. Here we use the parameters from ence of sulfate. Without sulfate, the carboxylic group is stable
the calibrated base case and compare predictions for five cases on the rock surfaces and the system remains oil-wet. With more

August 2015 SPE Journal 775

ID: jaganm Time: 19:17 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 776 Total Pages: 17

SO42– Solid-Surface Concentraion (μmol/m2)


(a)
0.5 days 5 days 10 days 20 days
mm mm mm mm
2.5
10 10 10 10
20 20 20 20 2
30 30 30 30 1.5
40 40 40 40
1
50 50 50 50
0.5
60 60 60 60
70 70 70 70 0
0 10 mm 0 10 mm 0 10 mm 0 10 mm

Water-Wet Fraction
(b)
0.5 days 5 days 10 days 20 days
mm mm mm mm 0.8
10 10 10 10
0.75
20 20 20 20
0.7
30 30 30 30
40 40 40 40 0.65

50 50 50 50 0.6
60 60 60 60 0.55
70 70 70 70
0 10 mm 0 10 mm 0 10 mm 0 10 mm

Oil Saturation
(c) 0.9
0.5 days 5 days 10 days 20 days
mm mm mm mm
0.8
10 10 10 10
20 20 20 20 0.7
30 30 30 30
0.6
40 40 40 40
50 50 50 50 0.5

60 60 60 60 0.4
70 70 70 70
0 10 mm 0 10 mm 0 10 mm 0 10 mm

Fig. 8—Spatial and temporal evolution of the SO22


4 surface concentration, water-wet fraction, and oil saturation for the imbibition
of SW. The corresponding oil-recovery curve is shown in Fig. 6c. The comparison of this figure with Fig. 7 explains the recovery
differences.

sulfate in the seawater, the ultimate oil recovery increased ent with multiple experimental observations (Zhang et al. 2006;
(SW0NaCl4SO4) by 5% OOIP and the rate of the oil produc- Fathi et al. 2011).
tion was also faster. These results demonstrate the importance A decrease in oil recovery is observed with increased aqueous
of sulfate in the wettability-alteration process and are consist- salinity, which is consistent with Fig. 7 and with experimental

Table 4—Parameters used in sensitivity analysis. The value for K6 was selected from the match of the base case because no such data
were found in the literature. Other values selected are within the range from the literature.

776 August 2015 SPE Journal

ID: jaganm Time: 19:18 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 777 Total Pages: 17

(a) (b)
0.8 0.8

0.6 0.6 log K6 = 4.4


Oil Recovery log K5 = 4.4

Oil Recovery
log K5 = 3.4 log K6 = 5.4
0.4 0.4

log K5 = 2.4
0.2 log K6 = 6.4
0.2

0 0
0 10 20 30 40 0 10 20 30 40
Time (days) Time (days)
(c) (d)
0.8 0.8

0.6 0.6 log K3 = –1.5


Oil Recovery

Oil Recovery
0.4 log K3 = –2.5
0.4
log K3 = –3.5
log K2 : –2.2
0.2 log K2 : –1.2 0.2

log K2 : –3.2
0 0
0 10 20 30 40 0 10 20 30 40
Time (days) Time (days)

Fig. 9—Oil-recovery curves for different reaction-equilibrium-constant parameters for Reactions 5 (a), 6 (b), 2 (c), and 3 (d). With
the same brine compositions, the significant difference in the oil recovery illustrates that the solid/water interface reactions (5 and
6) are very important. The specific values of these parameters can be found in Table 4.

observations (Fathi et al. 2011). Increasing Ca2þ concentration SW0NaCl, whereas the oil recovery for SW was predicted. There
has a negligible effect, suggesting that Ca2þ concentration is not a are two main differences between the model prediction and exper-
limiting quantity for SW0NaCl and there are abundant divalent imental data. First, the predicted oil recovery for SW is larger
cations in seawater. Similar marginal effects of Ca2þ were also than the experimental data in the first 5 days. There are multiple
observed in experiments (Fathi et al. 2011). reasons that could have caused the discrepancy. The lack of the
measured specific imbibition capillary pressure function for limit-
ing wettability cases could be one reason for the discrepancy.
Limitations and Potential Improvements. In the model valida- Another possible factor is numerical diffusion introduced by the
tion in Fig. 6, the oil-recovery curves were matched for FW and low-order finite-volume discretization. Numerical diffusion
causes additional mixing between the imbibing and connate fluids
and may overestimate the rates of wettability alteration. In addi-
0.8
tion, the ions were assumed to have the same diffusion coefficient.
However, small ions such as Naþ can diffuse faster than large
Chalk TSS = 0.6 x 109/g ions such as SO2 4 . Surface adsorption/desorption reactions could
0.6 also be dominated by kinetic effects, which violates the reaction-
Oil Recovery

equilibrium assumption (Bruemmer et al. 1988; Fuller et al.


1993). All these factors could lead to the mismatch of the oil-re-
0.4 Chalk TSS = 6 x 109/g covery rate in the first 5 days.
Second, the final oil-recovery factor for SW is underestimated.
Chalk TSS = 14.4 x 109/g
This can be because of the lack of accurate surface-chemistry
0.2 data. Reaction 6 used in this model was not validated by experi-
ments, but instead was tuned to provide a good match of the simu-
lated data to the experiments. Whether the acidic-oil-component
0 adsorption obeys the mass-action law is not clear. The diffusive-
0 10 20 30 40 layer model used to predict the effect of surface potential on car-
Time (days)
boxylic-group adsorption could also be a constraint of the model.
An independent measurement of the key reaction-equilibrium
Fig. 10—Oil recovery as a function of the total number of sur- constant K6 could allow a conclusive validation of the model.
face sites. A difference in the recovery rate is apparent at Also, the relative permeability and capillary pressure data specific
approximately Day 2. The ultimate recovery is the same, how- to the core experiments (at the initial wettability and final wett-
ever. The adsorbing capacity, represented by TSS, affects the ability) will lead to a complete validation of this model and better
recovery rate. understanding of the system.

August 2015 SPE Journal 777

ID: jaganm Time: 19:18 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 778 Total Pages: 17

0.8
1
Oil AN: 0.17
Oil AN: 0.17
0.6

Water-Wet Fraction
Oil Recovery

0.9
0.4 Oil AN: 1.90

Oil AN: 2.07 Oil AN: 1.90


0.8
0.2
Oil AN: 2.07

0 0.7
0 10 20 30 40 0 5 10 15
(a) Time (days) (b) Distance to Core Center (mm)

Fig. 11—(a) Oil recovery as a function of the AN. (b) Water-wet fraction profile along the radial direction at Day 10. Crude-oil acidity
affects the wettability alteration and oil recovery.

0.8
mm mm
D: 2.4x10–10 m2/s
0.6 10 10 0.02
Oil Recovery

20 20
D: 1.4x10–10 m2/s 0.015
Height

Height
0.4 30 30
D: 5.1 x10–11 m2/s 40 40 0.01

0.2 50 50
0.005
60 60
70 70 0
0 0 10 mm 0 10 mm
0 20 40 60 80
Radius Radius
(a) Time (days) (b) (c)

Fig. 12—(a) Predicted oil recovery with different values of diffusion coefficients. (b) The sulfate concentration in the aqueous
phase (mol/kg water) profile on Day 10 for diffusion coefficient 1:43310210 m2/s. (c) The sulfate concentration in the aqueous phase
(mol/kg water) profile on Day 10 for diffusion coefficient 0:51310211 m2/s. The oil-recovery rate (a) is very sensitive to the diffusion
coefficient because the diffusion process dominates the solute transport (b) in these experiments.

Conclusion waterflooding in mixed-wet carbonate reservoirs. The simulation


We developed a model to understand and predict wettability alter- results of spontaneous-imbibition experiments demonstrated that
ation and improved oil recovery (IOR) for chemically tuned our model can be used to describe the complex interactions
among different phases and species involved in chemically tuned
water IOR. The model captured the interplay among aqueous spe-
0.7 cies (Ca2þ, Mg2þ, SO2 4 ), crude-oil acidity, and solid-surface
SW0NaCl4SO4 properties. The coupled mixed-wettability and reactive-transport
0.6 model was capable of predicting the wettability-alteration effects
owing to chemical reactions. The oil-recovery curves predicted
0.5 SW0NaCl for different brine compositions showed good consistency with
experimental observations.
Oil Recovery

SW0NaCl4Ca
0.4 Our model improves the understanding of and provides a
powerful tool for chemically tuned waterflooding. The findings
SW include
0.3
• The surface-complexation reactions control the wettability-
0.2 alteration process. The chemically tuned brine should be
designed to promote oil-component desorption from the
SW0NaCl0SO4
0.1
rock. Moreover, because the oil recovery is sensitive to the
equilibrium constants, the specific reservoir condition
should be used in both experiments and simulation.
0 • There are two types of parameters. Type I controls the final
0 10 20 30 40
wettability and ultimate oil recovery, whereas Type II con-
Time (days) trols the rate of wettability alteration and rate of oil recov-
ery. Type I includes the reaction-equilibrium constants and
Fig. 13—Oil-recovery prediction for five chemically tuned crude-oil acid number. Type II includes total surface sites of
brines, where brine is the imbibing fluid at 110º C. A description
solid and the diffusion coefficient.
of the abbreviations is shown in Table 3. Generally, the oil re-
covery at 40 days increases with the sulfate concentration and • The concentrations of ionic species, ionic strength, and oil
decreases with the total ionic strength. This is because there acidity are important in the IOR process. The salinity should
are adequate Ca21 and Mg21 in seawater so that adding those not be the single factor that is used to assess the effective-
cations only has a marginal effect in such cases. ness of the chemically tuned water.

778 August 2015 SPE Journal

ID: jaganm Time: 19:18 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 779 Total Pages: 17

Nomenclature Sci. 84 (24 December): 218–241. http://dx.doi.org/10.1016/j.ces.


ai ¼ the activity of species i, dimensionless 2012.08.038.
a0i ¼ the ion-size parameter for Debye-Hückel model Anderson, W. G. 1987. Wettability Literature Survey Part 5: The Effects
A, b, B ¼ temperature-dependent parameters for the Debye- of Wettability on Relative Permeability. J Pet Technol 39 (11):
Hückel model 1453–1468. SPE-16323-PA. http://dx.doi.org/10.2118/16323-PA.
Aos ¼ area of the oil/solid interface, m2 Austad, T., Shariatpanahi, S. F., Strand, S., et al. 2012. Conditions for a
Aow ¼ area of the oil/solid interface, m2 Low-Salinity Enhanced Oil Recovery (EOR) Effect in Carbonate Oil
Aws ¼ area of the water/solid interface, m2 Reservoirs. Energ. Fuel. 26 (1): 569–575. http://dx.doi.org/10.1021/
Cio ¼ the concentration for the species io on the oil surface, ef201435g.
mol/m2 Austad, T., Strand, S., Madland, M. V., et al. 2008. Seawater in Chalk: An
Cis ¼ the concentration for the species is on the solid surface, EOR and Compaction Fluid. SPE Res Eval & Eng 11 (4): 648–654.
mol/m2 SPE-118431-PA. http://dx.doi.org/10.2118/118431-PA.
Ciw ¼ the concentration for the species iw in the aqueous Brady, P. V. and Krumhansl, J. L. 2012. A Surface Complexation Model
phase, mol/m3 of Oil–Brine–Sandstone Interfaces at 100 C: Low Salinity Waterflood-
D ¼ diffusion coefficient, m2/s ing. J. Pet. Sci. Eng. 81 (January): 171–176. http://dx.doi.org/10.1016/
F ¼ Faraday’s constant, 9.648104 C/mol j.petrol.2011.12.020.
Fp ¼ molar rate of the primary species p, mol/md Brady, P. V., Krumhansl, J. L. and Mariner, P. E. 2012. Surface Complex-
Fq ¼ molar rate of the secondary species q, mol/md ation Modeling for Improved Oil Recovery. Presented at the SPE
g ¼ gravity factor, m/s2 Improved Oil Recovery Symposium, Tulsa, Oklahoma, 14–18. SPE-
I ¼ the ionic strength of water, mol/kg water 153744-MS. http://dx.doi.org/10.2118/153744-MS.
k ¼ permeability, md Borgwardt, R. H. and Bruce, K. R. 1986. Effect of Specific Surface Area
kr ¼ relative permeability on the Reactivity of CaO with SO2. AIChE J. 32 (2): 239–246. http://
Kn;eq ¼ the equilibrium constant for the nth reaction, unitless dx.doi.org/10.1002/aic.690320210.
Mp ¼ molar density of the primary species p, mol/m3 Bruemmer, G. W., Gerth, J. and Tiller, K. G. 1988. Reaction Kinetics of
Mq ¼ molar density of the secondary species q, mol/m3 the Adsorption and Desorption of Nickel, Zinc and Cadmium by Goe-
Np ¼ the number of the primary species thite. I. Adsorption and Diffusion of Metals. J. Soil Sci. 39 (1): 37–52.
Nsec ¼ the number of secondary reactions http://dx.doi.org/10.1111/j.1365-2389.1988.tb01192.x.
Ntot ¼ the total number of species Buckley, J. 1994. Chemistry of the Crude Oil/Brine Interface. Proc., 3rd
P ¼ pressure, Pa International Symposium on Evaluation of Reservoir Wettability and
Pc ¼ the oil/water capillary pressure Its Effect on Oil Recovery, Laramie, Wyoming, 21–23 September,
R ¼ the gas constant, J/Kmol 33–38.
Ri ¼ reaction rate for the species i, mol/m3s Buckley, J. S. and Liu, Y. 1998. Some Mechanisms of Crude Oil/Brine/
S ¼ saturation, dimensionless Solid Interactions. J. Pet. Sci. Eng. 20 (3–4): 155–160. http://
T ¼ the absolute temperature, K dx.doi.org/10.1016/S0920-4105(98)00015-1
u ¼ fluid-volume flow rate, m3/d
~ Buckley, J. S., Bousseau, C. and Liu, Y. 1996. Wetting Alteration by Brine
zi ¼ the charge carried by the species i and Crude Oil: From Contact Angles to Cores. SPE J. 1 (3): 341–350.
a ¼ phase subscript, w representing water phase and o rep- SPE-30765-PA. http://dx.doi.org/10.2118/30765-PA.
resenting oil phase Cassie, A. B. D. 1948. Permeability to Water and Water Vapour of Tex-
ci ¼ the activity coefficient of species i, kg water/mol tiles and Other Fibrous Materials. Introductory Paper. Discuss. Fara-
cos ¼ the interfacial tension (IFT) of the oil/solid interface, day Soc. 3: 239–243. http://dx.doi.org/10.1039/DF9480300239.
dynes/cm Chapman, D. 1913. A Contribution to the Theory of Electrocapillarity.
cow ¼ the IFT of the oil/water interface, dynes/cm Philos. Mag. 25 (148): 475–481. http://dx.doi.org/10.1080/14786440
cws ¼ the IFT of the water/solid interface, dynes/cm 408634187.
e0 ¼ the dielectric constant of water, 55.3, dimensionless
Coats, K. H. 2000. A Note on IMPES and Some IMPES-Based Simulation
em ¼ the permittivity of free space, 8.854  1013 C
Models. SPE J. 5 (3): 245–251. SPE-65092-PA. http://dx.doi.org/
V 1 dm1
10.2118/65092-PA.
h ¼ the contact angle, degrees
l ¼ viscosity, poise Dang, C. T. Q., Nghiem, L. X., Chen, Z. J., et al. 2013. Modeling Low Sa-
li ¼ chemical potential of species i, J/mol linity Waterflooding: Ion Exchange Geochemistry and Wettability
pq ¼ the (q; p) entry in the stoichiometry matrix Alteration. Presented at SPE Annual Technical Conference and Exhi-
q ¼ density, kg/m3 bition, New Orleans, Louisiana, 30 September–2 October. SPE-
ro ¼ the charge density at the oil surface, C/m2 166447-MS. http://dx.doi.org/10.2118/166447-MS.
/ ¼ porosity, dimensionless Davis, J. A. and Kent, D. B. 1990. Surface Complexation Modeling in
wo ¼ the oil/water-interface-charge potential, mV Aqueous Geochemistry. In Mineral-Water Interface Geochemistry, ed.
ws ¼ the solid/water-interface-charge potential, mV M.F. Hochella Jr. and A. F. White. Rev. Mineral. 23: 177–260.
Davis, J. A., Meece, D. E., Kohler, M., et al. 2004. Approaches to Surface
Complexation Modeling of Uranium (VI) Adsorption on Aquifer Sedi-
Acknowledgments ments. Geochim. Cosmochim. Ac. 68 (18): 3621–3641. http://dx.doi.
The authors gratefully thank the member companies of the org/10.1016/j.gca.2004.03.003.
Enhanced Oil Recovery JIP in the EMS Energy Institute at Penn- Delshad, M., Lenhard, R. J., Oostrom, M., et al. 2003. A Mixed-Wet Hys-
sylvania State University at University Park, Pennsylvania, for teretic Relative Permeability and Capillary Pressure Model for Reser-
their financial support. voir Simulations. SPE Res Eval & Eng 6 (5): 328–334. SPE-86916-
PA. http://dx.doi.org/10.2118/86916-PA.
Delshad, M., Najafabadi, N. F., Anderson, et al. 2009. Modeling Wettabil-
References ity Alteration by Surfactants in Naturally Fractured Reservoirs. SPE
Alotaibi, M., Azmy, R. and Nasr-El-Din, H. 2010. Wettability Challenges Res Eval & Eng 12 (3): 361–370. SPE-100081-PA. http://dx.doi.org/
in Carbonate Reservoirs. Presented at the SPE Improved Oil Recovery 10.2118/100081-PA.
Symposium, Tulsa, Oklahoma, 24–28 April. SPE-129972-MS. http:// Drummond, C. and Israelachvili, J. 2004. Fundamental Studies of Crude
dx.doi.org/10.2118/129972-MS. Oil–Surface Water Interactions and its Relationship to Reservoir Wett-
Andersen, P., Evje, S., Madland, M. V., et al. 2012. A Geochemical Model ability. J. Pet. Sci. Eng. 45 (1–2): 61–81. http://dx.doi.org/10.1016/
for Interpretation of Chalk Core Flooding Experiments. Chem. Eng. j.petrol.2004.04.007.

August 2015 SPE Journal 779

ID: jaganm Time: 19:18 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 780 Total Pages: 17

Dzombak, D. and Morel, F. 1990. Surface Complexation Modeling: Environ. Sci. Technol. 45 (23): 9959–9966. http://dx.doi.org/10.1021/
Hydrous Ferric Oxide. Hoboken, New Jersey: John Wiley & Sons. es201111y.
Evje, S. and Hiorth, A. 2011. A Model for Interpretation of Brine- Li, L., Steefel, C. I., Kowalsky, M. B., et al. 2010. Effects of Physical and
Dependent Spontaneous Imbibition Experiments. Adv. Water Geochemical Heterogeneities on Mineral Transformation and Biomass
Resour. 34 (12): 1627–1642. http://dx.doi.org/10.1016/j.advwatres. Accumulation During Biostimulation Experiments at Rifle, Colorado.
2011.09.003. J. Contam. Hydrol. 112 (1): 45–63. http://dx.doi.org/10.1016/j.jcon
Fathi, S. J., Austad, T. and Strand, S. 2010. “Smart Water” as a Wettability hyd.2009.10.006.
Modifier in Chalk: The Effect of Salinity and Ionic Composition. Lichtner, P. C. 1985. Continuum Model for Simultaneous Chemical Reac-
Energ. Fuel. 24 (4): 2514–2519. http://dx.doi.org/10.1021/ef901304m. tions and Mass Transport in Hydrothermal Systems. Geochim. Cosmo-
Fathi, S. J., Austad, T. and Strand, S. 2011. Water-Based Enhanced Oil chim. Ac. 49 (3): 779–800. http://dx.doi.org/10.1016/0016-7037(85)
Recovery (EOR) by “Smart Water”: Optimal Ionic Composition for 90172-3.
EOR in Carbonatese. Energ. Fuel. 25 (11): 5173–5179. http:// Lichtner, P. C. 1996. Continuum formulation of multicomponent-multi-
dx.doi.org/10.1021/ef201019k. phase reactive transport. In Reactive Transport in Porous Media,
Fung, L. S.-K., Hiebert, A. D. and Nghiem, L. X. 1992. Reservoir Simula- Reviews in Mineralogy, 34, 1–81.
tion with a Control-Volume Finite-element Method. SPE Res Eval & McGuire, P. L., Chatham, J. R., Paskvan, F. K., et al. 2005. Low Salinity
Eng 7 (3): 349–357. SPE-21224-PA. http://dx.doi.org/10.2118/21224- Oil Recovery: an Exciting New EOR Opportunity for Alaska’s North
PA. Slope. Presented at the SPE Western Regional Meeting, Irvine, Cali-
Fuller, C. C., Davis, J. A. and Waychunas, G. A. 1993. Surface Chemistry fornia, 30 March–1 April. SPE-93903-MS. http://dx.doi.org/10.2118/
of Ferrihydrite: Part 2. Kinetics of Arsenate Adsorption and Coprecipi- 93903-MS.
tation. Geochim. Cosmochim. Ac. 57 (10): 2271–2282. http:// Megawati, M., Hiorth, A. and Madland, M. V. 2013. The Impact of Sur-
dx.doi.org/10.1016/0016-7037(93)90568-H. face Charge on the Mechanical Behavior of High-Porosity Chalk.
Gibbs, J. 1948. The Collected Works of J. Willard Gibbs, Vol. 1. New Ha- Rock Mechanics and Rock Engineering 46 (5): 1073–1090.
ven, Connecticut: Yale University Press. Morrow, N. R. 1990. Wettability and its Effect on Oil Recovery. J Pet
Gouy, G. 1910. On the Formation of Electrical Charges at the Surface of Technol 42 (12): 1476–1484. SPE-21621-PA. http://dx.doi.org/
an Electrolyte. J. Physique 9: 457–469. http://dx.doi.org/10.1051/ 10.2118/21621-PA.
jphystap:019100090045700. Nasralla, R. A. and Nasr-El-Din, H. A. 2012. Double-Layer Expansion: Is
Gupta, R. and Mohanty, K.K. 2011. Wettability Alteration Mechanism It A Primary Mechanism of Improved Oil Recovery by Low-Salinity
for Oil Recovery from Fractured Carbonate Rocks. Transport Waterflooding. Presented at the SPE Improved Oil Recovery Sympo-
Porous Med. 87 (2): 635–652. http://dx.doi.org/10.1007/s11242- sium, Tulsa, Oklahoma, 14–18 April. SPE-154334-MS. http://
010-9706-5. dx.doi.org/10.2118/154334-MS.
Hallenbeck, L. D., Sylte, J. E., Ebbs, D. J., et al. 1991. Implementation of O’Carroll, D. M., Abriola, L. M., Polityka, C. A. et al. 2005. Prediction of
the Ekofisk Field Waterflood. SPE Form Eval 6 (3): 284–290. SPE- two-phase capillary pressuresaturation relationships in fractional wett-
19838-PA. http://dx.doi.org/10.2118/19838-PA. ability systems. Journal of Contaminant Hydrology 77 (4): 247–270.
Helgeson, H. C., Brown, T. H., Nigrini, A., et al. 1970. Calculation of http://dx.doi.org/10.1016/j.jconhyd.2005.01.004.
Mass Transfer in Geochemical Processes Involving Aqueous Solu- Owens, W. W. and Archer, D. L. 1971. The Effect of Rock Wettability on
tions. Geochim. Cosmochim. Ac. 34 (5): 569–592. http://dx.doi.org/ Oil-Water Relative Permeability Relationships. J Pet Technol 23 (7):
10.1016/0016-7037(70)90017-7. 873–878. SPE-3034-PA. http://dx.doi.org/10.2118/3034-PA.
Helland, J. O. and Skjaeveland, S. M. 2006. Physically Based Capillary PennSim Toolkit. 2013. Gas Flooding Joint Industry Project, EMS
Pressure Correlation for Mixed-Wet Reservoirs From a Bundle-of- Energy Institute, Pennsylvania State University, University Park,
Tubes Model. SPE J. 11 (2): 171–180. SPE-89428-PA. http:// Pennsylvania.
dx.doi.org/10.2118/89428-PA. Peters, E. J. 2012. Advanced Petrophysics: Solutions, Vol. 3. Austin,
Hill, D. 1984. Diffusion Coefficients of Nitrate, Chloride, Sulphate, and Texas: Live Oak Book Company.
Water in Cracked and Uncracked Chalk. Journal of Soil Science 35 Puntervold, T. and Austad, T. 2008. Injection of Seawater and Mixtures
(1): 27–33. http://dx.doi.org/10.1111/j.1365-2389.1984.tb00256.x. with Produced Water into North Sea Chalk Formation: Impact of
Hiorth, A., Cathles, L. M. and Madland, M. V. 2010. The Impact of Pore Fluid–Rock Interactions on Wettability and Scale Formation. J. Pet.
Water Chemistry on Carbonate Surface Charge and Oil Wettability. Sci. Eng. 63 (1–4): 23–33. http://dx.doi.org/10.1016/j.petrol.2008.
Transport Porous Med. 85 (1):1–21. http://dx.doi.org/10.1007/s11242- 07.010.
010-9543-6. Puntervold, T., Strand, S. and Austad, T. 2009. Coinjection of Seawater
Hirasaki, G. and Zhang, D. L. 2004. Surface Chemistry of Oil Recovery and Produced Water to Improve Oil Recovery From Fractured North
From Fractured, Oil-Wet Carbonate Formations. SPE J. 9 (2): Sea Chalk Oil Reservoirs. Energ. Fuel. 23 (5): 2527–2536. http://
151–162. SPE-88365-PA. http://dx.doi.org/10.2118/88365-PA. dx.doi.org/10.1021/ef801023u.
Hjelmeland, O. S. and Larrondo, L. E. 1986. Experimental Investigation Puntervold, T., Strand, S. and Austad, T. 2007. Waterflooding of Carbon-
of the Effects of Temperature Pressure and Crude Oil Composition on ate Reservoirs: Effects of a Model Base and Natural Crude Oil Bases
Interfacial Properties. SPE Res Eval & Eng 1 (4): 321–328. SPE- on Chalk Wettability. Energ. Fuel. 28 (5): 1606–1616. http://
12124-PA. http://dx.doi.org/10.2118/12124-PA. dx.doi.org/10.1021/ef060624b.
Høgnesen, E. J., Standnes, D. C. and Austad, T. 2006. Experimental and RezaeiDoust, A., Puntervold, T., Strand, S., et al. 2009. Smart Water as
Numerical Investigation of High Temperature Imbibition into Prefer- Wettability Modifier in Carbonate and Sandstone: A Discussion of
ential Oil-Wet Chalk. J. Pet. Sci. Eng. 53 (1): 100–112. http:// Similarities/Differences in the Chemical Mechanisms. Energ. Fuel. 23
dx.doi.org/10.1016/j.petrol.2006.04.002. (9): 4479–4485. http://dx.doi.org/10.1021/ef900185q.
Jerauld, G. R., Webb, K. J., Lin, C.-Y., et al. 2008. Modeling Low-Salinity Robertson, E. P. 2007. Low-Salinity Waterflooding to Improve Oil Recov-
Waterflooding. SPE Res Eval & Eng 11 (6): 1000–1012. http:// ery-Historical Field Evidence. Presented at the SPE Annual Technical
dx.doi.org/10.2118/102239-PA. Conference and Exhibition, Anaheim, California, 11–14 November.
Lager, A., Webb, K. J., Black, C. J. J., et al. 2006. Low Salinity Oil SPE-109965-MS. http://dx.doi.org/10.2118/109965-MS.
Recovery–An Experimental Investigation. Oral presentation given at Rosen, M. J. and Kunjappu, J. T. 2012. Surfactants and Interfacial Phe-
the International Symposium of the Society of Core Analysts, Trond- nomena. Hoboken, New Jersey: John Wiley & Sons.
heim, Norway, 12–16 September. Steefel, C. I. and Lasaga, A. C. 1994. A Coupled Model for Transport of
Langmuir, D. 1997. Aqueous Environmental Geochemistry. Upper Saddle Multiple Chemical Species and Kinetic Precipitation/Dissolution
River, New Jersey: Prentice Hall. Reactions with Application to Reactive Flow in Single Phase Hydro-
Li, L., Gawande, N., Kowalsky, M. B., et al. 2011. Physicochemical Het- thermal Systems. Am. J. Sci. 294 (5): 529–592. http://dx.doi.org/
erogeneity Controls on Uranium Bioreduction Rates at the Field Scale. 10.2475/ajs.294.5.529.

780 August 2015 SPE Journal

ID: jaganm Time: 19:18 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 781 Total Pages: 17

Steefel, C. I., DePaolo, D. J. and Lichtner, P. C. 2005. Reactive Transport Oil Recovery Symposium, Tulsa, Oklahoma, 14–18 April. SPE-
Modeling: An Essential Tool and a New Research Approach for the 154076-MS. http://dx.doi.org/10.2118/154076-MS
Earth Sciences. Earth Planet Sc. Lett. 240 (3): 539–558. http:// Yousef, A., Al-Saleh, S. and Al-Jawfi, M. S. 2012b. The Impact of the
dx.doi.org/10.1016/j.epsl.2005.09.017. Injection Water Chemistry on Oil Recovery from Carbonate Reser-
Strand, S., Høgnesen, E. J. and Austad, T. 2006a. Wettability Alteration of voirs. Presented at the SPE EOR Conference at Oil and Gas West
Carbonates—Effects of Potential Determining Ions (Ca2þ and SO2 4 )
Asia, Muscat, Oman, 16–18 April. SPE-154077-MS. http://dx.doi.org/
and Temperature. Colloid. Surface. A 275 (1): 1–10. http://dx.doi.org/ 10.2118/154077-MS.
10.1016/j.colsurfa.2005.10.061. Yousef, A.A., Liu, J. S., Blanchard, G. W., et al. 2012c. SmartWaterflood-
Strand, S., Puntervold, T. and Austad, T. 2008. Effect of Temperature on ing: Industry’s First Field Test in Carbonate Reservoirs. Presented at
Enhanced Oil Recovery From Mixed-Wet Chalk Cores by Spontane- the SPE Annual Technical Conference and Exhibition, San Antonio,
ous Imbibition and Forced Displacement Using Seawater. Energ. Fuel. Texas, 8–10 October. SPE-159526-MS. http://dx.doi.org/10.2118/
22 (5): 3222–3225. http://dx.doi.org/10.1021/ef800244v. 159526-MS.
Strand, S., Standnes, D. C. and Austad, T. 2006b. New Wettability Test Yu, L., Evje, S., Kleppe, H., et al. 2009. Spontaneous Imbibition of Sea-
for Chalk Based on Chromatographic Separation of SCN and SO2 water Into Preferentially Oil-Wet Chalk Cores—Experiments And
4 . J.
Pet. Sci. Eng. 52 (1): 187–197. http://dx.doi.org/10.1016/j.petrol.2006. Simulations. J. Pet. Sci. Eng. 66 (3): 171–179. http://dx.doi.org/
03.021. 10.1016/j.petrol.2009.02.008
Zhang, P. and Austad, T. 2005a. Waterflooding in Chalk: Relationship
Strand, S., Standnes, D. C. and Austad, T. 2003. Spontaneous Imbibition
Between Oil Recovery, New Wettability Index, Brine Composition
of Aqueous Surfactant Solutions Into Neutral to Oil-Wet Carbonate
And Cationic Wettability Modifier. Presented at the SPE Europec/
Cores: Effects of Brine Salinity and Composition. Energ. Fuel. 17 (5):
EAGE Annual Conference, Madrid, Spain, 13–16 June. SPE-94209-
1133–1144. http://dx.doi.org/10.1021/ef030051s.
MS. http://dx.doi.org/10.2118/94209-MS.
Tang, G.-Q. and Morrow, N. R. 1999. Influence of Brine Composition and Zhang, P. and Austad, T. 2005b. The Relative Effects of Acid Number
Fines Migration on Crude Oil/Brine/Rock Interactions and Oil Recov- and Temperature on Chalk Wettability. Presented at SPE International
ery. J. Pet. Sci. Eng. 24 (2–4): 99–111. http://dx.doi.org/10.1016/ Symposium on Oilfield Chemistry, The Woodlands, Texas, 2–4 Febru-
S0920-4105(99)00034-0. ary. SPE-92999-MS. http://dx.doi.org/10.2118/92999-MS.
Ustohal, P., Stauffer, F. and Dracos, T. 1998. Measurement and Modeling Zhang, P. and Sparks, D. L. 1990. Kinetics and Mechanisms of Sulfate
of Hydraulic Characteristics of Unsaturated Porous Media with Mixed Adsorption/Desorption on Goethite Using Pressure-Jump Relaxation.
Wettability. J. Contam. Hydrol. 33 (1–2): 5–37. http://dx.doi.org/ Soil Science Society of America Journal 54 (5): 1266–1273. http://
10.1016/S0169-7722(98)00063-1. dx.doi.org/10.2136/sssaj1990.036159950054000500011x.
Villalobos, M., Trotz, M. A. and Leckie, J. O. 2001. Surface Complexa- Zhang, P., Tweheyo, M. T. and Austad, T. 2006. Wettability Alteration
tion Modeling of Carbonate Effects on the Adsorption of Cr(VI), and Improved Oil Recovery In Chalk: The Effect Of Calcium in The
Pb(II), and U(VI) on Goethite. Environ. Sci. Technol. 35 (19): Presence Of Sulfate. Energ. Fuel. 20 (5): 2056–2062. http://dx.doi.org/
3849–3856. http://dx.doi.org/10.1021/ES001748K. 10.1021/ef0600816.
Walter, A. L., Frind, E. O., Blowes, D. W., et al. 1994. Modeling of Multi- Zhang, P., Tweheyo, M. T. and Austad, T. 2007. Wettability Alteration
component Reactive Transport in Groundwater: 1. Model Develop- And Improved Oil Recovery By Spontaneous Imbibition of Seawater
ment and Evaluation. Water Resour. Res. 30 (11): 3137–3148. http:// Into Chalk: Impact of The Potential Determining Ions Ca2þ, Mg2þ,
dx.doi.org/10.1029/94WR00955. and SO24 . Colloid. Surface. A 301 (1–3): 199–208. http://dx.doi.org/
Webb, K. J., Black, C. J. J. and Al-Ajeel, H. 2004. Low Salinity Oil 10.1016/j.colsurfa.2006.12.058.
Recovery–Log-Inject-Log. Presented at the SPE/DOE Symposium on Zysset, A., Stauffer, F. and Dracos, T. 1994. Modeling of Reactive
Improved Oil Recovery, Tulsa, Oklahoma, 17–21 April. SPE-89379- Groundwater Transport Governed by Biodegradation. Water Resour.
MS. http://dx.doi.org/10.2118/89379-MS. Res. 30 (8): 2423–2434. http://dx.doi.org/10.1029/94WR01045.
Webb, K. J., Black, C. J. J. and Tjetland, G. 2005. A Laboratory Study
Investigating Methods for Improving Oil Recovery in Carbonates. Pre-
sented at the International Petroleum Technology Conference, Doha, Appendix A: Mass Conservation for Multiphase
Qatar, 21–23 November. IPTC-10506-MS. http://dx.doi.org/10.2523/ Reactive Transport System
10506-MS. This appendix derives the multiphase reactive-transport equation,
Wolery, T. J., Jackson, K. J., Bourcier, W.L., et al. 1990. Current Status of Eq. 5. For the aqueous species, the mass-conservation equation
the EQ3/6 Software Package for Geochemical Modeling. In Chemical for the primary species iw is
Modeling of Aqueous Systems II, Vol. 146, ed. D. C. Melchior and R.
L. Bassett, Chapter 8, 104–116. Washington, DC: American Chemical @
Society. http://dx.doi.org/10.1021/bk-1990-0416.ch008. ð/Sw qw Ciw Þ þ r  ½Ciw qw~
u w  DrðSw qw Ciw Þ
@t
Yeh, G.-T. and Tripathi, V. S. 1991. A Model for Simulating Transport of
Reactive Multispecies Components: Model Development and Demon-
¼ Req kinetic
iw þ Riw ; iw ¼ 1; …; nw : . . . . . . . . . . . . . ðA-1Þ
stration. Water Resour. Res. 27 (12): 3075–3094. http://dx.doi.org/
10.1029/91WR02028. For the water/solid-interface species iws and the oil/solid-inter-
Yildiz, H. O. and Morrow, N. R. 1996. Effect of Brine Composition on face species ios , the mass-conservation equation is
Recovery of Moutray Crude Oil by Waterflooding. J. Pet. Sci. Eng. 14
@
(3): 159–168. http://dx.doi.org/10.1016/0920-4105(95)00041-0. ½ð1  /Þqs Aws Ciws  ¼ Req kinetic
iws þ Riws ; iws ¼ 1; …; nws ;
Yousef, A. A., Al-Saleh, S., Al-Kaabi, A. U., et al. 2010. Laboratory @t
Investigation of Novel Oil Recovery Method for Carbonate Reservoirs.                    ðA-2Þ
Presented at the Canadian Unconventional Resources and International @
Petroleum Conference, Calgary, Alberta, Canada, 19–21 October. ½ð1  /Þqs Aos Cios  ¼ Req kinetic
ios þ Rios ; ios ¼ 1; …; nos :
@t
SPE-137634-MS. http://dx.doi.org/10.2118/137634-MS.
Yousef, A., Al-Saleh, S. and Al-Jawfi, M. S. 2011. Smart WaterFlooding
                   ðA-3Þ
for Carbonate Reservoirs: Salinity and Role of Ions. Presented at the
SPE Middle East Oil and Gas Show and Conference, Manama, Bah- For the oil/water-interface species iow , the mass-conservation
rain, 25–28 September. SPE-141082-MS. http://dx.doi.org/10.2118/ equation is
141082-MS.
@
Yousef, A. A., Al-Saleh, S. and Al-Jawfi, M. S. 2012a. Improved/ ½ð1  /Þqs Aow Ciow  ¼ Req kinetic
iow þ Riow ; iow ¼ 1; …; now ;
@t
Enhanced Oil Recovery from Carbonate Reservoirs by Tuning Injec-
tion Water Salinity and Ionic Content. Presented at the SPE Improved                    ðA-4Þ

August 2015 SPE Journal 781

ID: jaganm Time: 19:18 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 782 Total Pages: 17

where Ciw is the concentration of species iw in aqueous phase where all the reactions are assumed to be at equilibrium and the
(mol/kg water); Ciws is the concentration of species iws on water/ kinetic rates vanish. We followed the typical reactive-transport
solid interface (mol/m2); Cios is the concentration of species ios on formulation to partition the reactive species into primary and sec-
oil/solid surface (mol/m2); Ciow is the concentration of species iow ondary species. The rate of production of a primary species p
on oil/water interface (mol/m2); D is the diffusion/dispersion from the equilibrium reactions can be written as the total rates of
coefficient (m2/s); Aws is the water/solid-interface area (m2/g production of secondary species:
solid); Aos is the oil/solid-interface area (m2/g solid); and Aow is XNsec
the oil/brine-interface area (m2/g solid). In Eq. A-1, the first term Req
p ¼  qp Rq ;
is for mass accumulation; the second term is a transport term, q¼1
which includes convection and diffusion; and the right-hand side where qp represents the (q, p) entry of the stoichiometry coeffi-
is the reaction term. The reaction rates are classified into equilib- cient matrix. Here the primary species p or secondary species q
rium-controlled reaction rates and kinetically controlled reaction can represent aqueous species, solid-surface species, or species on
rates, denoted as Req and Rkinetic (mol/s-kg water). These two the oil/water interface. We can eliminate the reaction rates from
terms can be the summation of multiple reaction-rate terms, Eqs. A-1, A-5, and A-6 for Eq. 5:
depending on the number of reactions that involve this species.
! !
Eqs. A-2, A-3, and A-4 do not have the transport term because the @ X
Nsec X
Nsec
interfaces are not mobile in our model. Mp þ qp Mq þ r  Fp þ qp Fq ¼ 0;
The interface between brine and oil contains surface-active @t q¼1 q¼1
sites as the polar components in crude-oil hydrate, whereas the p ¼ 1; …; np :                       ð5Þ
interface between brine and solid contains solid-surface sites that
contain polar ions (Buckley 1994; Buckley and Liu 1998; Hira-
saki and Zhang 2004). The interface between the solid and oil Appendix B: Relation Between Surface
phase is regarded as the part of the solid surface that is fully occu- Concentration and Contact Angle
pied with organic ion-containing sites. At the solid surface, the One crucial step in our model is to relate surface geochemistry to
competition between the oil and water species for the solid sur- surface-contact angles. That relation was built by estimating inter-
face determines the proportion of the solid surface that is occu- facial tension (IFT) from a Gibbs isotherm. The contact angle was
pied by water and oil. This occupancy of solid-surface sites then estimated from Eq. 7. The change in IFT can be quantified
determines the wettability. by the Gibbs adsorption equation (Gibbs 1948),
The total solid-surface area (Aos þ Aws ) represents the total X
available surface area and is a constant calculated on the basis of dcws ¼  Ci dli;ws ; . . . . . . . . . . . . . . . . . . . . . . ðB-1Þ
the specific surface area (SSA) of the rock. The total oil-surface i
area (Aos þ Aow ) may vary depending on oil saturation. We define where dcws is the differential change of the water and surface IFT,
the solid-surface concentration (Ci;s ) as Ci is the surface-excess concentration of species i (mol/m2), and
8A dli;ws is the differential change of the chemical potential of spe-
>
<
ws
Ciws ; when i ¼ iws : cies i on the surface (mJ/mol). For surface-active species, the sur-
Ci;s ¼ AAs face-excess concentration Ci can be considered to be equal to the
>
: os Cios ; when i ¼ ios : actual surface concentration without significant error (Rosen and
As Kunjappu 2012); namely, Ci ¼ ½i, where ½i is the surface concen-
tration of species i in mol/m2. The chemical potential of surface-
The oil/water-surface concentration can be defined similarly. species i can be calculated from
Eqs. A-2 through A-4 can be rewritten as the mass conservation
on the total solid surface and total oil surface: li;ws ¼ l0i;ws þ RTlnai;ws ¼ l0i;ws þ RTlnðci ½iÞ;
@
½ð1  /Þqs As Cis  ¼ Req kinetic
is þ Ris ; . . . . . . . . . . . . ðA-5Þ where l0i;ws is the chemical potential of surface-species i at a refer-
@t ence state. If we assume the activity-coefficient ci has a weak de-
@ pendence on the surface concentration ½i, then
½ð1  /Þqs Ao Cio  ¼ Req
i þ Ri
kinetic
; . . . . . . . . . . . . ðA-6Þ
@t RT
dli;ws ¼ d½RTlnðci ½iÞ ¼ RTdðln½ci  þ ln½iÞ ¼ d½i:
where Cis is the solid-surface concentration that includes the oil/ ½i
solid and water/solid species; Cio is the oil-surface concentration
that includes the oil/water species; and As and Ao are the total sur- Substitution of dli;ws into Eq. B-1 gives
face areas of solid and oil, respectively. In this research, As is the X X
RT
total surface area and is calculated on the basis of the specific sur- dcws ¼  ½i d½i ¼ RT i d½i: . . . . . . . . . . ðB-2Þ
face area of the carbonate rocks, whereas Ao is a linear function of
i ½i
oil saturation.
Eq. B-2 indicates that the surface tension is linear with respect to
We define the molar density of species i as
the surface concentration. The derivation is also valid for an oil/
8 solid interface. In the low-salinity-flooding scenario, the change
< /Sw qw Ciw ; for i ¼ iw
Mi ¼ ð1  /Þqs As Cis ; for i ¼ is ; in cow is not considered significant (Zhang et al. 2006). Therefore,
:   X
ð1  /Þqs Ao Cio ; for i ¼ io cos  cws dcos  dcws RT i d½i
dcoshw ¼ d ¼ ¼ :
and the molar-flow rate of species i as cow cow cow
8                    ðB-3Þ
u w  DrðSw Ciw Þ; for i ¼ iw
< Ci;aq qw~ Eq. B-3 indicates that for an ideal system, cosh is a linear function
Fi ¼ 0; for i ¼ is : of the surface concentration of sulfate. A linear interpolation can
: be used when we know extreme values of the contact angles.
0; for i ¼ io
With h ¼ h0 for fresh water and h ¼ h1 for a sulfate-surface con-
From these definitions, we can write a general equation for the centration > Ca; total, at surface concentration of > CaSO 4 the
transport of chemical species: effective contact angle is estimated by Eq. 8:
@ ½> CaSO 4
Mi þ r  Fi ¼ Req
i ;
coshw ¼ cosh0 þ ðcosh1  cosh0 Þ: . . . . . . ð8Þ
@t ½> Ca; total

782 August 2015 SPE Journal

ID: jaganm Time: 19:18 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014


J170966 DOI: 10.2118/170966-PA Date: 6-August-15 Stage: Page: 783 Total Pages: 17

Changhe Qiao is a PhD degree candidate in the Department and phase behavior, unconventional gas engineering, multi-
of Mathematics at the Pennsylvania State University. Qiao’s phase flow in porous media, and well testing. He received the
research interest includes compositional simulation, reactive- SPE Ferguson Medal in 1993 and served as coexecutive editor
transport models, fast and robust linear solvers, and parallel for SPE Reservoir Evaluation and Engineering from 2002 to
computing. He holds a bachelor’s degree in computational 2004. In 2009, Johns was awarded the SPE Distinguished Mem-
mathematics from Peking University, China. ber Award and in 2013 the SPE Faculty Pipeline Award. He is
currently director of the Enhanced Oil Recovery Consortium
Li Li is an assistant professor in the Department of Energy and and codirector of the Unconventional Natural Resources Con-
Mineral Engineering at Pennsylvania State University. She has sortium in the EMS Energy Institute at Pennsylvania State
worked for more than 10 years in understanding and predicting University. Johns holds a bachelor’s degree in electrical engi-
physical-, chemical-, and biological-process coupling in spa- neering from Northwestern University and master’s and PhD
tially heterogeneous subsurface. In addition to low-salinity flood- degrees in petroleum engineering from Stanford University.
ing, Li has worked on a range of topics at the water/energy
nexus, including geological carbon sequestration, microbe Jinchao Xu is the Verne M. Willaman Professor of Mathematics,
enhanced oil recovery, reservoir souring, and bioremediation. the Francis R. and Helen M. Pentz Professor of Science, and
Director of the Center for Computational Mathematics and
Russell T. Johns is the Chair of the Petroleum and Natural Gas
Applications at Pennsylvania State University. His research
Engineering Program within the Department of Energy and
Mineral Engineering at Pennsylvania State University. He also interests are in the design, analysis, and application of numeri-
holds the Victor and Anna Mae Beghini Professorship of Petro- cal methods, especially multilevel and adaptive finite-element
leum and Natural Gas Engineering and the CMG Foundation methods, for systems of partial-differential equations and
Chair in fluid behavior and rock interactions. Before his current problems with direct applications to physics, biology, and en-
position, Johns served on the petroleum engineering faculty gineering. Xu was a plenary speaker at the International Con-
at the University of Texas at Austin from 1995 to 2010. He also gress for Industrial and Applied Mathematics in 2007 and an
has 9 years of industrial experience as a petrophysical engi- invited speaker at the International Congress for Mathemati-
neer with Shell Oil and as a consulting engineer for Colenco cians in 2010. He is a fellow of the Society for Industrial and
Power Consulting, Baden, Switzerland. Johns has more than Applied Mathematics and the American Mathematical Soci-
200 publications in enhanced oil recovery, thermodynamics ety in the inaugural class.

August 2015 SPE Journal 783

ID: jaganm Time: 19:18 I Path: S:/J###/Vol00000/150014/Comp/APPFile/SA-J###150014

You might also like