Anda di halaman 1dari 12

Article

pubs.acs.org/EF

Modeling Low-Salinity Waterooding in Chalk and Limestone


Reservoirs
Changhe Qiao,* Russell Johns, and Li Li
Pennsylvania State University, University Park, Pennsylvania 16802, United States

ABSTRACT: The injection of low-salinity brines can improve oil recovery in carbonate reservoirs by changing the rock
wettability from being more oil-wet to being more water-wet. Existing models use an empirical dependence of wettability based
on variables including equivalent salinity and ionic strength. We recently developed a process-based model that mechanistically
includes the geochemical interactions between crude oil, brine, and the chalk surface that alter rock wettability. In this research,
we extend the previous model by including mineral dissolution reactions, therefore enabling the modeling of low-salinity ooding
in chalks and limestone cores with and without anhydrite, which is considered to be a key factor in controlling the extent of
improved oil recovery (IOR). We examine the role of mineralogy by including surface complexation, aqueous reactions, and
dissolution/precipitation of calcite and anhydrite in the extended model. These reactions, coupled with the equations of
multiphase ow and transport, are solved simultaneously using an in-house simulator, PennSim. Relative permeability functions
and residual oil saturation during ooding are adjusted dynamically according to the concentration of oil acids attached to the
mineral surface. Core ooding experiments from the Stevns Klint (SK) chalk and a Middle Eastern carbonate with a small
volume fraction of anhydrite are used to tune the reaction network and predict recovery. Simulation results agree with the
observed euent concentrations of SO42, Ca2+, and Mg2+ reported from chromatographic wettability tests and the measured
recoveries under diering compositions in chalk and limestone cores. For the SK chalk without anhydrite, lower Na+ and Cl
concentrations under constant SO42 conditions leads to IOR by as much as 6% OOIP. Lower salinity alone, however, does not
lead to IOR in limestones without anhydrite. Instead, anhydrite dissolution provides a natural source of sulfate and increases oil
recovery by 5% when injecting diluted formation water. Simulations of two-dimensional (2D) ve-spot patterns using tuned
reaction networks demonstrated that IORs from 5% to 20% OOIP can be obtained after two pore volumes are injected. These
IORs are greatly dependent on the aqueous chemistry of the injected uid and sweep. The results highlight the critical
importance of understanding the mineralogy and including a mechanistic reaction model in the simulation of low-salinity water
oods.

INTRODUCTION
Low-salinity water (LSW) has been reported in the literature to
with chemically tuned water conrmed wettability alteration by
incremental sulfate adsorption onto the rock surface.6,12
improve oil recovery signicantly in sandstone and carbonate Wettability alteration was also conrmed in contact angle
core experiments.14 Here, we focus on the widely used term measurements on a at plane carbonate surface, where smaller
low-salinity ooding, although both experimental data from contact angles were reported with modied seawater.2
Austad et al.1,57 and our model8 suggest that improved oil Furthermore, capillary pressure reduction was observed in
recovery (IOR) is not solely caused by low salinity (total spontaneous imbibition tests when sulfate-containing seawater
dissolved salt). Instead, it is the specic composition of the was injected, compared to sulfate-free water.13 Shifts in relative
injection waterin particular, the concentrations of the permeability and decreases in residual oil saturation were
wettability-altering species (SO42, Ca2+, and Mg2+)that demonstrated in core ooding experiments.10
matters. The recoveries in core oods and imbibition tests vary Wettability alteration could be caused by mineral dissolution,
greatly, depending on the aqueous chemistry used, the pore surface adsorption/desorption, and surface potential reduc-
volumes injected (PVI), and the carbonate mineralogy.6,9 tion.7,14,15 Austad et al.1 proposed that, for the Stevns Klint
Improved recoveries were 5% to 40% OOIP for chalk outcrop, the concentrations of potential determining ions
spontaneous imbibition tests in the Stevns Klint chalk and (PDIs) such as Mg2+, Ca2+ and SO42 greatly aected the oil
5% to 20% OOIP in core ooding experiments in both the recovery in spontaneous imbibition tests. They explained that,
Stevns Klint chalk and a Middle Eastern limestone.6,2,10 Diluted
for mixed wet surfaces, both SO42 and carboxylic groups in the
formation water without sulfate was found to improve oil
oil adsorb onto the rock surface, altering the surface charge
recovery by 5% OOIP.5 Diluted seawater was reported to
improve oil recovery by 17% to 19% OOIP in a limestone core from positive to negative. The negatively charged solid surface
that contained anhydrite.2 A single well tracer test at eld scale then attracts cations such as Mg2+ and Ca2+ to react with the
also exhibited decreased residual oil saturation by seven adsorbed carboxylic group, which releases the carboxylic group
saturation units from the injection of diluted seawater.11
One of the primary mechanisms proposed for IOR by LSW Received: October 18, 2015
ooding is wettability alteration from a more oil-wet state to a Revised: January 11, 2016
more water-wet state.1,10 Chromatographic wettability tests

XXXX American Chemical Society A DOI: 10.1021/acs.energyfuels.5b02456


Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

and causes the surface to become more water-wet. The notion limestones with anhydrite. We rst introduce the multiphase
that carboxylic groups leave the chalk surface was inferred from reactive transport equations, and then we describe the
water vapor adsorption isotherms, where SO42 reduced the wettability alteration model and the mechanisms of altered
surface concentrations of adsorbed fatty acids.16 Spontaneous wettability and oil recovery in dierent cores. Lastly, we use the
imbibition experiments with varied brine compositions support new model to demonstrate the potential of LSW ooding at the
such chemical mechanisms.6,7,9,12 In addition to SO42, BO32 eld pattern scale.
and PO43 were also found to improve oil recovery.17
Experimental studies have revealed that the brine composi-
tion sucient for enhanced oil recovery in Stevns Klint chalk
METHODOLOGY
We consider isothermal recovery processes with immiscible oil/water
was not necessarily eective for other carbonate minerals. ow. Geochemical reactions occur in the aqueous phase and at the
Instead, diluted formation water and seawater were found to crude oil/brine-mineral interfaces. Desorption of the carboxylic group
improve oil recovery the most for limestone and dolomites.2,18 from the rock surface leads to more water-wet conditions, which
Romanuka et al.18 performed spontaneous imbibition tests for further changes the relative permeability and residual oil saturation
based on measured curves at two wettability end points. The major
Stevns Klint chalk, limestone, and dolomite cores and extension from our previous model8 includes the addition of calcite
concluded that eective EOR uids are signicantly dependent and anhydrite dissolution/precipitation reactions, surface complex-
on the rock mineralogy. Screening Amott imbibition experi- ation with the surface site >CO3 , and additional aqueous complexation
ments showed that additional SO42 did not lead to IOR for such as CaCl+ and NaSO4. [Note that the prex > denotes the
limestone and dolomite, as well as for the Stevns Klint chalk species at the solid surface.] In this section, we describe the governing
with diluted formation water. One possible reason for an equations that couple the ow, transport and reaction system, and the
increase in oil recovery for sulfate-free injection uids in inputs to setup a simulation.
limestone is the increased sulfate concentration from Governing Equations. The equation set describes the mass
dissolution of anhydrite in situ.5 In addition, SO42 and Mg2+ conservation of oil/water and chemical species, Darcy ow, and
reaction kinetics and equilibrium. The physical meanings and units of
were reported to improve oil recovery incrementally in core the symbols are given in the Nomenclature section.
ooding experiments for limestone.19 Immiscible Oil/Water Flow Equations. The mass conservation
The above-mentioned experimental evidences have shown equations for immiscible oil and water phases are as follows:
signicant promise of LSW for IOR; however, at the same time,

it has indicated the complexity of geochemical reactions (S ) + ( u ) = 0 = o, w
(1)
involved in low-salinity ooding in dierent types of carbonate t
formations. Currently, there is no theory to quantify and Darcys law governs the ow rate of each phase,
explain many of these dierences. Various modeling eorts kr
have tried to capture the wettability alteration process. u = K(P gZ) = o, w
(2)
Wettability alteration has been modeled by interpolating
capillary pressure, relative permeability, and residual oil The subscript w represents the water phase and subscript o
saturation between two wetting states.20,21 Empirical relation- represents the oil phase. Capillary pressure relates the pressure of oil
ships between wettability and water chemistry parameters were and water phases:
used including salinity, ionic strength, or total dissolved Pcow = Po Pw (3)
salts.20,22 Contact angles were also used to represent surface
wettability.22 Although these relationships are valid under the The saturation relation completes the set of equations:
given experimental conditions, they do not take into account So + Sw = 1 (4)
complex surface reactions mechanistically and therefore cannot
be used to predict wettability alteration under conditions that The primary unknowns for the multiphase ow system are Po and Sw.
Reactive Transport Equations. The mass conservation equation for
dier from where the empirical relationships are derived.23 the primary species p is
Thus, a mechanistic model is needed to understand the
Nsec Nsec
dominant processes and to accurately predict wettability under
varying mineralogical. (M p + qpMq) + (Fp + qpFq) = 0 p = 1 , ..., Np
t q=1 q=1
Qiao et al.8 developed a mechanistic model that coupled
(5)
multiphase ow and transport using a detailed surface and
aqueous multicomponent reaction network including adsorp- where the rst term is the accumulation of total moles and the second
tion/desorption that captures the competitive interactions term is the total molar ux of a primary component p, the chemical
building blocks of the reaction systems. The equations are based on
among oil, brine, and the surface of the Stevns Klint chalk.
the stoichiometric relationship among the species participating in
Natural carbonate reservoirs, however, often contain minerals reactions and reaction kinetics. The derivation of the reactive transport
including anhydrite, dolomite, and quartz. These minerals could equations are given in Qiao et al.8 More details of reactive transport
potentially aect surface reactions among rock, oil, and water, equations can also be found in Yeh and Tripathi,25 Lichtner,28 Steefel
and therefore inuence the oil recovery. For example, Austad et and Lasaga,26 and Walter et al.27 These equations solve for the
al.5 reported that anhydrite can play an important role in concentrations of primary species in the aqueous phase and at the
wettability alteration. phase interfaces. Secondary species are solved using laws of mass
In this paper, we extend the multiphase reactive transport action based on the concentrations of primary species. Reactive
model by Qiao et al.8 to include limestone surface complexation transport models have been widely used in subsurface biogeochemistry
in the past decades;2831 however, they are relatively new in
and mineral dissolution/precipitation reactions. This extends applications of interest to the oil and gas industry.24,3234
the model in our in-house simulator PennSim24 to provide a Multiphase Reaction Network. Reactions are either kinetically
unied model for low-salinity ooding in mineralogically controlled or at equilibrium, depending on the typical rates of the
dierent carbonate reservoirs. The model was calibrated using specic geochemical reactions. According to the geochemical
core experimental data from the Stevns Klint chalk and convention, the reactions in the aqueous phase and surface

B DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

Table 1. Multiphase Reaction Networka


number reaction log Keq(25 C) log Keq(110 C)
Oil/Water Interface Reactions
O1 COOH COO + H+ 5.00 5.00
O2 +
COOCa COO + Ca 2+ 1.20 1.20
O3 +
COOMg COO + Mg 2+ 1.00 1.30
Solid (Calcite)/Water Interface Reactions
C1 > CaOH + H+ >CaOH+2 11.80 11.80
C2 > CaOH+2 + SO24 > CaSO4 + H 2O 2.10 3.25

C3 > CaOH+2 + CO32 > CaCO3 + H 2O 6.00 6.00

C4 > CO3H > CO3 + H+ 5.10 5.10


C5 > CO3Ca + > CO3 + Ca 2 + 2.50 3.40

C6 > CO3Mg + > CO3 + Mg 2 + 2.50 3.40

CalciteWater/OilWater Interface Reaction


CO1 > CaOH+2 ( COO) >CaOH+2 + COO 5.90
Aqueous-Phase Reactions
A1 H 2O H+ + OH 13.99 12.24
A2 HCO3 H + +
CO32 10.33 10.27

A3 H 2CO3 H+ + HCO3 6.35 6.43


A4 MgSO4 Mg 2+
+ SO24 2.37 2.46

A5 NaSO4 Na + + SO24 0.70 0.77

A6 CaSO4 Ca 2+
+ SO24 2.30 2.61

A7 +
CaCl Ca 2+
+ Cl 0.70 0.54
A8 +
MgCl Mg 2+
+ Cl 0.15 0.74
Mineral Dissolution/Precipitation Reactions
M1 CaCO3(s) + H Ca 2 + + HCO3
+ 1.85 0.58

M2 CaSO4 (s) Ca 2 + + SO24 4.36 5.34


a
Keq values for oil/water interface reactions are taken from Brady and Krumhansl;47 Keq for calcite surface reactions are taken from Pokrovsky and
Schott42 and Brady et al.;40 Keq for aqueous and mineral reactions are taken from Wolery.37 KCO1 was xed at 5.9 obtained from reproducing initial
wettability of Fathi et al.6 The kinetic rate constant for reaction M1 is 106.19 mol/(m2 s);59 the kinetic rate constant for reaction M2 is 102.95 mol/
(m2 s).60 The new additions in the reactions, compared to the previous model,8 include reactions C3C6, A4A8, M1, and M2.

complexation on the rock and oil surfaces typically occur fast and are ai = iCi
assumed to be at equilibrium.35 In contrast, mineral dissolution and
precipitation are relatively slow and modeled with reaction kinetics.36 Activity coecients are determined using an extended DebyeHuckel
The reaction network, along with the thermodynamic and kinetic model.38
constants at 25 and 110 C, are listed in Table 1. The primary species High Na+ and Cl concentrations reduce surface adsorption in two
include aqueous ions HCO3 , Cl, SO2 2+ 2+ +
4 , SCN , Mg , Ca , Na , K ,
+
ways. First, higher concentrations lead to greater ionic strength and
H+, and surface species > CaOH+2 , > CO3 and COO, where the smaller activity coecients of aqueous species (and, therefore, less
prexes > and represent the species at the solid surface and oil surface adsorption). Second, Na+ and Cl form aqueous complexes
surface, respectively. The extension from Qiao et al.8 includes calcite with SO2 2+ 2+
4 , Mg , and Ca through reactions A5, A7, and A8, which
surface reactions (reactions C4C6), aqueous reactions (reactions reduces the corresponding free-ion concentration available for surface
A4A8) and mineral dissolution/precipitation reactions (reactions M1 adsorption. Therefore, reducing Na+ and Cl concentrations can
and M2). The equilibrium constants for aqueous and mineral promote surface adsorption, which causes wettability alteration.
dissolution reactions are taken from the EQ3/6 thermodynamic Mineral Surface Reactions. We considered the surface dissolution/
database.37 precipitation of calcite and anhydrite (reactions M1 and M2 in Table
Aqueous Complexation Reactions in Water. The aqueous species
1). The mineral reaction rates are written based on the transition state
include free ions HCO3 , Cl, SO2 2+ 2+ + +
4 , SCN , Mg , Ca , Na , K , H ,
+
theory (TST) rate laws.39 For example, calcite dissolution rate is
CO23 , OH , and H2CO3 and complex ions MgSO4, NaSO4 , CaSO4,
expressed as follows:
CaCl+, and MgCl+. An example, the equilibrium constant of reaction
A4 is
E aCa 2 +a HCO3
R CaCO3(s) = A CaCO3(s)k H+ exp a a H+1
Keqa H +
a Mg 2 +aSO24 RT
Keq,A4 =
a MgSO
4
The last term in the parentheses in the above equation indicates how
where the activities of the aqueous species (ai) are calculated by far away the reaction is from equilibrium. If the reaction is at

C DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

equilibrium, the rate is zero. The Damkohler number (D) for this wet, as observed in low-salinity waterooding. If the reaction goes
reaction is calculated as from right to left, the system becomes more oil-wet, as observed
during the aging of crude oil. This reaction is reversible over the time
R CaCO3(s) scale of core ooding, as shown in a recent experimental paper by
Da ,CaCO3(s) =
vCCa 2 +,total Hopkins et al.49 at a temperature of 50 C. Reaction CO1 happens fast
at high temperature but is much slower at room temperature.1,5052
where v is the ow velocity and C Ca 2+,total is the total Ca 2+ This reaction is assumed to be reversible at high temperature (110
concentration. The Damkohler number determines if the reaction is C). For simulation of the chromatographic wettability test at 25 C,
modeled using an equilibrium law or a kinetic rate law. Calcite we assumed irreversibility during the duration of the experiments (3
dissolution provides a source for Ca2+ species and buers pH. The h). The equilibrium constant of this reaction is not known in the
calculated Damkohler number for anhydrite dissolution is more than literature, so we used this parameter as a tuning parameter. In this
100 in the experimental cases studied later in this paper, indicating research, the value of KCO1 was xed at 5.9, as obtained from
much faster dissolution rates than transport. Therefore, the anhydrite reproducing the initial wettability of Fathi et al.,6 and this value was
dissolution reaction is modeled at equilibrium. Anhydrite dissolution is used for all other simulations. Therefore, the surface concentration of
an important source or sink for sulfate and calcium. >CaOH2+(COO) is a wettability indicator: the higher this
In addition to the mineral dissolution reactions, we also consider concentration, the more oil-wet the solid surface. The wettability
the surface adsorption/desorption reactions on a calcite surface. indicator () is calculated from
Surface complexation models are typically used to describe the surface c c ww
reactions.4043 A calcite surface contains sites >CaOH and >CO3H =
that can adsorb aqueous ions and carboxylic groups and form surface cow c ww (6)
species >CaOH2+, >CaSO4 , >CaCO3, >CO3, >CO3Ca+, and
where c, c ww , and c ow are the surface concentrations of
>CO3Mg+.44 In this research, two models are used to describe the
>CaOH+2 (COO) at the current state, the end-point water-wet
surface reactions: the electrical diusive layer (EDL) and the
state, and the oil-wet state, respectively. Values of cww and cow are
nonelectrical diusive layer (NEDL) model (the latter of which is
obtained by speciation calculation at the initial and injection water
also called the constant capacitance model).43,45 We used the relatively
compositions.
simple-to-implement NEDL model when the change of surface
For relative permeability values, we use the following linear
potential does not have a signicant eect. The EDL model is used
interpolation model:
when there are signicant surface potential changes, because of surface
adsorption. Thus, the EDL model also aects the surface adsorption * = (1 )k rw,ww
k rw * *
+ k rw,ow
equilibrium. In the EDL model, the activity of an aqueous species is
corrected through a Boltzmann factor, so the equilibrium constant for * = (1 )k ro,ww
k ro * + k ro,ow
* (7)
reaction C2 is expressed as follows:
* , krw,ow
where kro,ow * , kro,ww
* , and krw,ww
* are relative permeability values at
2F

K C2 =
( )
exp RTs [>CaSO4 ] the end-point oil-wet (ow) and end-point water-wet states (ww).
These end-point states are not required to be completely oil-wet or
[>CaOH+2 ]aSO24 water-wet, but ideally should be measured under initial reservoir
conditions (mixed wet state) and at the most water-wet state possible.
Here [>CaSO4 ] and [>CaOH+2 ] are the surface concentrations (in The BrooksCorey model is used to calculate relative permeability at
units of mol/m2). The surface potential (s) is calculated from the any arbitrary saturation between the end points,53
GouyChapman theory,
* (S*)n w
k rw = k rw
F
s = 80mIRT sinh s
2RT *(1 S*)no
k ro = k ro
In the NEDL model, s is considered constant. The equilibrium where the normalized water saturation (S*) is calculated as
constant for reaction C2 can be written then as
Sw Swi
S* =
[>CaSO4 ] 1 Swi Sor
K C2 =
[>CaOH+2 ]aSO24
Here, Swi is the irreducible water saturation and Sor is the residual oil
here the constant surface potential term is lumped into KC2. A saturation. Alteration of the wettability can change the contact angle,
sensitivity analysis in Qiao et al.46 demonstrated the strong which reduces the capillary number and then the residual oil
dependency of the oil recovery on the equilibrium constants. For saturation. Residual oil saturation has been shown to change during
aqueous reactions, the Keq values are well-documented for a large low salinity waterooding, as shown in the works of Nasralla et al.10
range of temperature. For surface complexation, the Keq values are and Morrow et al.54 Therefore, Sor is a function of wettability, as
empirical parameters related to intrinsic thermodynamic constants.45 follows:
The solid surface site densities for >CaOH and >CO3H are xed at 4.1 Sor = (1 )Sorww + Sorow
mol/m2, based on a previous simulation study.40
Oil/Water Interface Reactions. The active species on the oil/water Figure 1 shows two examples of relative permeability curves
interface include COOH, COO, COOCa+, and COOMg+. measured or calculated under end-point wettability conditions.
The NH4 oil surface sites are not included here because COO Relative permeability end points under other wettability conditions
dominates the interaction between oil and water surfaces.40 Oil surface are determined using eq 7. Ideally, the relative permeability curves
complexation reactions are described by either NEDL or EDL in a should be measured at the end-point wettability states. Capillary
similar way as mineral surface reactions. Temperature eects for pressure is also important when the capillary number is large.
reactions on the oil surface are not considered to be as signicant.47 However, most of the coreooding experiments do not report the
The surface oil site total concentrations are xed at 6 mol/m2.48 relative permeability and capillary pressure curves. In this paper,
Interfacial Interactions and Wettability Alteration. The calcite- relative permeability curves at the end points were determined by the
water/oil interface reaction is represented by reaction CO1. The best t to the measured oil recoveries, because data were not available.
species >CaOH+2 (COO) is the adsorbed carboxylic group on the The capillary pressure is assumed to be zero, since the relative
calcite surface. When the reaction goes from left to right, permeability curve is tuned, and we prefer not to have more tuning
>CaOH+2 (COO) desorbs and the system becomes more water- parameters. The capillary pressure eects are specically studied for

D DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

(1) Tune kro, krw, and Sor at the original wetting state to match oil
recovery with injection of formation water where wettability
alteration does not occur.
(2) Tune Sor at the end-point water-wet case to match ultimate oil
recovery.
(3) Perform speciation calculations given the formation water
composition to obtain cow. Namely, solve the nonlinear
equation systems that describe the equilibrium condition of
all reactions and conservation of the total concentration as the
formation water. Perform speciation calculation given the LSW
that leads to the most water wettability to obtain cww. Select
Figure 1. Relative permeability curves at the end-point wetting states either the NEDL or the EDL model so that cow > cww.
used in the simulations of the core ooding experiments in (A) Fathi (4) Perform simulations using the end-point parameters obtained
et al.,6 with 100% volume fraction of calcite, and (B) Yousef et al.,2 in steps 1, 2, and 3.
with 3% anhydrite. In both cases, Swi remains constant when the Reaction equilibrium constants are xed for simulation of all
wettability changes. The Sor values decrease more in panel (B) than in experiments. We conduct one-dimensional (1-D) simulations of core
panel (A). The curves for panel (A) show signicantly less changes in ooding experiments in dierent carbonate rocks listed in Table 2.
the relative permeability curves, compared to panel (B), because the The 1-D simulation domains were discretized uniformly into 100 grid
core in Fathi et al.6 is initially more water-wet than the core in Yousef blocks, which is the minimum number of grid blocks needed to
et al.2 reproduce the tracer breakthrough curves in the chromatographic
wettability test. Oil and water viscosities were xed at 1.0 and 0.476 cP,
respectively. We also used the reactions and tuned parameters from
the 1-D core oods to model two-dimensional (2-D) ve-spot
spontaneous imbibition in a previous publication.8 Experimental patterns.
studies also shows that the change in Swi is negligible for high-salinity
water and LSW.55 The initial water saturation was assumed to be
constant. The water saturation remains high after alteration of the
wettability. Even Swi changes, the oil recovery will not be aected.
RESULTS AND DISCUSSION
Stevns Klint Chalk Outcrop (No Anhydrite). The Stevns
Simulation Cases. Table 2 lists the mineralogical compositions Klint chalk is comprised of 55% (volume) calcite with 45% pore
and parameters for the limestone and chalk cores. The Stevns Klint space. The wettability can be altered by injection of ions that
chalk in Fathi et al.6 and the Middle Eastern limestone in Strand et causes the oil acids to desorb from the surface, changing the
al.56 contain pure calcite, with 100% volume fraction in the solid phase. rock to a more water-wet state. Austad et al.1 demonstrated this
In contrast, the limestone cores from Austad et al.5 and Yousef et al.2 in imbibition tests where a core plug was immersed by aqueous
contain anhydrite. In these experiments, kr curves were not measured. solution containing SO42. Forced displacement experiments
A practical workow for simulation of core ooding experiments (core oods) also show that chemically tuned seawater
where relative permeability curves are not measured is given as follows. injection leads to increased oil recovery in the same cores.6

Table 2. Input Parameters for Simulation of Core Flooding Experimentsa


Fathi et al.,6 Yousef et al.,2
SCC#1 Strand et al.,56 Core 2-21 Austad et al.,5 Core 5B 1st core ooding
core material Stevns Klint chalk Middle Eastern Limestone from the aqueous zone of an oil Middle Eastern limestone
limestone reservoir
mineral volume fraction 55% calcite 75% calcite 79% calcite, 3% anhydriteb 71% calcite, 3% anhydritec
porosity 0.45 0.25 0.18 0.25
permeability 1.0 mD 2.7 mD 1.2 mD 39.6 mD
specic surface area
calcited 1.00 m2/g 0.29 m2/g 0.29 m2/g 0.29 m2/g
anhydrite 1.0 m2/g 1.0 m2/g
dimensions
length, L 70.0 mm 49.0 mm 81.0 mm 40.6 mm
diameter, D 38.1 mm 37.8 mm 38.0 mm 38.0 mm
back pressure 145 psi 101.5 psi 145 psi 1000 psi
temperature 25 C 70 C 110 C 100 C
injection rate 0.20 mL/min 0.10 mL/min 0.01 mL/min 1.00 mL/min
surface complexation model NEDL NEDL NEDL EDL
cwwe 0.40 mol/m2 1.35 mol/m2 0.03 mol/m2
cowe 2.1 mol/m2 1.67 mol/m2 0.17 mol/m2
diusion/dispersion 0.0042 ft2/day 0.0020 ft2/day 0.0020 ft2/day 0.0020 ft2/day
coecientf

a
Values were taken from the reference listed in the rst row, unless otherwise specied. bThe 3% anhydrite was speculated, according to Austad et
al.5 cYousef et al.2 reported that the crushed powder is composed of 80% calcite, 13% dolomite, 6% anhydrite, and <1% quartz. The composition
used in simulations was calculated for the porous media and by lumping dolomite with calcite without dierentiating between the two. dThe specic
surface area for limestone was taken from Shariatpanahi et al.61 eThe end-point concentrations were obtained from speciation calculations with water
compositions at end-point oil-wet and water-wet states. fThe diusion/dispersion coecient was determined from tracer euent concentration
curves.

E DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

Table 3. Synthetic Brine Compositions Used in Experiments and Simulationsa


FWb SWb SW0NaClb SW0NaCl4SO4b SW1/2Mc NaCl-Mc FWd seawatere
ion concentration (mol/kg water)
HCO3 0.009 0.002 0.002 0.002 0.002 0.000 0.003 0.002
Cl 1.070 0.525 0.126 0.126 0.574 0.556 3.643 0.907
SO42 0 0.024 0.024 0.096 0.012 0.013 0 0.044
SCN 0 0 0 0 0.012
Mg2+ 0.008 0.045 0.045 0.045 0.045 0.013 0.076 0.088
Ca2+ 0.029 0.013 0.013 0.013 0.013 0.013 0.437 0.016
Na+ 1.000 0.450 0.050 0.194 0.475 0.504 2.620 0.796
K+ 0.005 0.010 0.010 0.010 0.022
pHf 8.4 8.4 8.4 8.4 8.4 8.4 8.4 8.4
total ionic strength 1.116 0.658 0.258 0.474 0.682 0.608 4.159 1.149
a
Legend: FW, formation water; SW, seawater; SW0NaCl, seawater depleted in NaCl; SW0NaCl4SO4, SW0NaCl with four levels of sulfate; SW1/
2M, seawater that contains SCN and half the amount of SO42 in SW; and NaCl-M, NaCl solution with equal amounts of Ca2+, Mg2+, and SCN.
b
Data taken from Fathi et al.6 cData taken from Strand et al.56 dData taken from Austad et al.5 eData taken from Yousef et al.2 fThe pH was not given
in the original references; here, we used a pH value of 8.4, following the pH in the chemically tuned water in the work of Zhang et al.9

Chromatographic Wettability Test. In chromatographic We included all reactions in the simulations, except reaction
wettability tests, seawater with dierent compositions (SW1 in CO1, because, at room temperature, the oil surface reactions
Table 3) is injected into a formation-water-saturated core and a were assumed to be inactive. Two simulations were performed
core initially at residual oil saturation (here, Sor = 0.25). The to reproduce the experimental breakthrough curves6 with and
purpose of chromatographic wettability tests is to quantify the without residual oil. The cores were lled with formation water
adsorption of sulfate. Figure 2A shows the breakthrough curves initially. Figure 2B shows the predicted Ca2+ and Mg2+
concentrations. Because the injected seawaterwater contains
more Mg2+ and less Ca2+ than the formation water does, more
Mg2+ adsorbed on the calcite surface, replacing adsorbed Ca2+
as the seawater contacted the rock surface, causing the Ca2+
concentration peak near 1.0 PVI. This is consistent with
experimental observations.6,9 Both concentrations eventually
returned to the injection concentrations.
Figure 2C shows the BTCs for cores at a residual oil
saturation of 0.25. The surface area that is in direct contact with
the aqueous phase is smaller than at the saturated conditions
without oil, leading to a smaller dierence between the SCN
and SO42 curves. The SCN and SO42 breakthrough earlier
than those in Figure 2A, because of the immobile residual oil
and less available pore space for ow and less surface area for
adsorption. The predicted Mg2+ and Ca2+ concentrations in
Figure 2D are similar to those depicted in Figure 2B, except
with an earlier and a smaller Ca2+ peak. The match of the BTCs
with and without oil shows the validity of the model and
parameters relevant to the physical processes when no oil
reactions are included.
Forced Displacement. Forced displacement experiments
Figure 2. Breakthrough curves for seawater injection into water-wet
were performed with successive injection of formation water
cores at room temperature: (A) SCN and SO42 breakthrough curves (FW), seawater (SW), and seawater depleted in NaCl
(BTCs) from experiments6 and simulations without oil, (B) Ca2+ and (SW0NaCl).6 The initial water saturation is 0.08.
Mg2+ BTCs from simulations without oil, (C) SCN and SO42 BTCs Simulation studies are performed in two ways. Simulation A
from experiments6 and simulations in the presence of residual oil, and injected FW, SW, and SW0NaCl successively in Table 3; in
(D) Ca2+ and Mg2+ BTCs from simulations with oil. C0 is the injection simulation B, NaCl solution was injected with the same total
concentration of the corresponding species. molality of FW, SW, and SW0NaCl. As shown in Figure 3A,
simulation A closely reproduced the experimental data. The oil
in a reference water-wet core without oil. The diusion/ recovery for simulation B, however, indicates that without
dispersion coecients (0.0042 ft2/day) were determined by SO42, Ca2+, and Mg2+, oil recovery was the same, even though
reproducing the breakthrough curves (BTCs) of the non- the injection water has increasingly lower salinity. This indicates
reactive tracer SCN. The area between the BTCs for SCN that, for the Stevns Klint chalk, it is the chemical composition
and SO42 quanties the amount of adsorbed SO42, as of the injected water that matters, not salinity. This is very
dierent from core ooding in limestones, where lower salinity
discussed in Zhang et al.,9 which essentially measures the has been shown to improve oil recovery.2,10,19,52
amount of surface sites available and the water-rock contact Figure 3B shows the predicted BTCs of total SO42,
surface area. >CaSO4, and >CaOH2(COO), as well as the pH for
F DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

Figure 3. (A) Oil recovery from experiments and simulations for


Figure 4. Breakthrough curves (BTCs) of chromatographic wettability
forced displacements with successive injection of FW, SW, and
tests in limestone cores:56 (A) SCN and SO24 at 25 C and (B)
SW0NaCl. The experimental data are taken from Fathi et al.6 (B) The
SCN, Ca2+, and Mg2+ at 70 C.
dimensionless concentration of total SO 4 2 , >CaSO 4 , and
>CaOH2(COO) for simulation A. C0 = 0.024 mol/kg water for
SO42, and 4.1 mol/m2 for surface species, >CaSO4, and
>CaOH2(COO). used to describe sulfate adsorption in limestone and in SK
chalk.
Surface Adsorption Test for Ca2+ and Mg2+. A NaCl
solution with equal amounts of Ca2+, Mg2+ and SCN (NaCl-
simulation A. During the FW injection, >CaOH2(COO) M, see Table 3) was injected to test the calcite surface
maintained a high concentration while no sulfate sorption adsorption of Ca2+ and Mg2+ in a similar chromatographic
occurred (>CaSO4 = 0). During the SW injection, the euent wettability test.56 As shown in Figure 4B, the breakthrough
concentrations and oil recovery did not change much until the times of Ca2+ and Mg2+ are longer than SCN, indicating that
seawater reached the outlet at 3.0 PVI, when the total Ca2+ and Mg2+are adsorbed on the surface. The adsorption
concentration of SO42 increased with increasing >CaSO4 and reactions are independent from SO42 adsorption and the
decreasing >CaOH2(COO). The high anity of sulfate to the carboxylic group, as the adsorption is observed without SO42
calcite surface, indicated by higher surface concentrations of and oil acids. Figure 4B shows the comparison of experimental
>CaSO4 than >CaOH2(COO), leads to increased oil and predicted BTCs for SCN, Ca2+, and Mg2+ at 70 C. The
recovery between 3.0 PVI and 4.0 PVI. From 4.0 PVI to 6.0 BTCs were matched with tuned KC5 and KC6 (2.1 for both
PVI, SW0NaCl was injected with the same SO42 concentration reactions), because of the increased temperature using the
as in the SW case, but without NaCl. With the decrease in ionic NEDL model, while other reaction parameters are the same as
strength, SO42 activity increased so that the surface that in the simulation of Fathi et al.6
concentration of >CaSO4 increased further while the aqueous The two simulations of the chromatographic wettability tests
total SO42 decreased. With continued SO42 injection, the indicated no signicant dierence between chalk and limestone,
total concentration eventually increased back to the inlet regarding the surface reactivity behaviors of Ca2+, Mg2+, and
concentration at 5.0 PVI. The increased sulfate adsorption SO42. The dierence in adsorption is a result of dierent
reduces >CaOH2(COO), which causes the observed change porosity and specic surface area. That is, we can use the same
in wettability toward a more water-wet state and the increased reaction network listed in Table 1 to describe the surface
in oil recovery. The predicted SO42 and >CaOH2(COO) complexations on the Stevns Klint chalk6 and the Middle
concentrations are consistent with the chemical mechanism Eastern limestone surfaces.56
proposed by Austad et al.1 The pH of the euent solution Limestone with Anhydrite. Anhydrite dissolution was
changes between 6.5 and 9.5 and is responsive to the changes in reported in some low-salinity core oods and was regarded as
injection water composition. Euent pH increased during the the cause of the alteration in wettability when the injection
SW slug and decreased during the SW0NaCl slug. After 3 uids do not contain sulfate. Here, we examine our model
PVI, the pH change mirrored the change in >CaSO4, indicating results for two cases.
the dominant control of SO42 sorption on the pH of the Forced Displacement in Limestone Core with Anhydrite
system. from Austad et al.5 Diluted formation water was found to
Limestone without Anhydrite from Strand et al.56 The improve oil recovery in limestone cores.5,2,10 Austad et al.5
selected limestone core was composed of 75% calcite and 25% observed tertiary IORs of 5% OOIP in limestone rocks with
pore space by volume. Thus, most of the reactions are the same diluted formation water containing no sulfate. Sulfate was
as those for the SK chalk. The limestone has a lower specic detected in the euent when the core was ooded with
surface area (0.29 m2/g), compared to that of chalk (1.00 m2/ deionized water, conrming the presence of anhydrite. The
g). increased SO42 concentration from dissolved anhydrite likely
Chromatographic Wettability Tests. To explore the anity led to the wettability alteration and increased IOR. To test this
of SO42 toward the limestone surface, a chromatographic hypothesis, we simulated the forced displacement experiments
wettability test for a completely water-wet core is simulated. In in Austad et al.,5 where formation water (FW) and 100-times-
the test, seawater that contains SCN and half of SO42 of SW diluted formation water (100 diluted FW) were injected
(SW1/2M) was injected (Table 3). A smaller diusion/ successively.
dispersion coecient (0.002 ft2/day) than the SK chalk was Figure 5A shows that simulation A with anhydrite dissolution
obtained by reproducing the BTC of SCN. As shown in reproduced the incremental oil recovery with 100 diluted FW.
Figure 4A, the predicted SO42 concentration agrees well with In simulation B, where anhydrite dissolution was removed,
the concentration data. The similarity between Figure 4A and Figure 5A indicates that the diluted FW does not lead to IOR.
Figure 2A indicates that the same equilibrium constants can be Figure 5B shows the predicted total euent concentration of
G DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

(BTCs) of Na + , SO 4 2 , and the concentration of


>CaOH2(COO) and >CaSO4. Stepwise decreases in Na+
concentration are due to the slug injection of diluted seawater.
The concentration of >CaOH2(COO) also decreased
stepwise, demonstrating the change in wettability to more
water-wet states. Because the concentration of >CaSO4 also
decreased, the reason for the decrease in >CaOH2(COO) is
not due to the competition by SO2 4 adsorption. A lower ionic
strength leads to a smaller solid surface potential and less
adsorption of the negatively charged carboxylic group.
Figure 5. (A) Oil recovery for forced displacements in cores with Comparison of Chalk and Limestones. Using the same
anhydrite dissolution (simulation A) and without anhydrite dissolution input parameters for the previously discussed cases, we
(simulation B). (B) Breakthrough curves of Ca2+, total Ca2+, SO42, compare predicted oil recoveries for the dierent cores with
and total SO42 from simulation. the same injection water composition. The core oods for
dierent mineralogies were modeled with the same parameters
in NEDL as shown in Figures 3A and 5A; core oods for
Ca2+ and SO42 for simulation A. Upon the injection of 100 dierent mineralogies with the same parameters in EDL were
diluted FW, Ca2+ concentrations decreased while SO42 modeled as shown in Figure 6A.
concentrations increased. Since the injection water does not Figure 7A compares the oil recovery of chalk6 from limestone
contain SO42, the total SO42 concentration indicates how A (Austad et al.5) and limestone B (Yousef et al.2), both
much anhydrite is dissolved. As shown in Figure 5B, total
SO42 did not change much when the injection uid
composition was changed at 6.0 PVI. This is because the
lower Ca2+ concentration in the injection water increased
anhydrite dissolution while a smaller ionic strength decreased
anhydrite dissolution. The net result is that anhydrite
dissolution remained almost the same as that indicated by the
almost-constant total SO42 concentration. The concentration
of free SO42, however, increased by more than one order of
magnitude, because Na+ and Ca2+ concentrations decreased
with the 100 diluted FW, so that less aqueous complexation
of NaSO4 and CaSO4(aq) were formed. Thus, additional free
SO42 was available to adsorb onto the solid surface to replace
adsorbed oil acids, leading to an increased oil recovery of 5%
OOIP.
Forced Displacement in Composite Limestone with
Anhydrite in Yousef et al.2 Yousef et al.2 observed 20%
OOIP incremental oil recovery with successive injection of
seawater, 2 diluted seawater, and 10 diluted seawater into
composite limestone cores. Anhydrite dissolution was veried
in their experiments using nuclear magnetic resonance (NMR)
measurements. The cores were originally oil-wet. The NEDL
model failed to predict any alteration of wettability in our rst Figure 7. (A) Oil recovery for chalk, limestone A, and limestone B
attempt. An EDL model, however, reproduced the data, when seawater, 2 diluted seawater, and 10 diluted seawater were
injected. (B) The ratio of the breakthrough concentration of adsorbed
indicating the importance of including the surface potential carboxylic acids >CaOH2(COO), with respect to its initial
changes in simulations with brines of dierent ionic strength. concentration for each case in panel (A). (C) Oil recovery for chalk,
Figure 6A shows that the model reproduced oil recovery limestone A, and limestone B; formation water (FW) and 100
data. Figure 6B shows the predicted breakthrough curves diluted FW were injected. (D) The ratio of the breakthrough
concentration of adsorbed carboxylic acids >CaOH2(COO), with
respect to its initial concentration at the condition of panel (C).
Simulation input parameters are the same as those in Figure 3A for
chalk, Figure 5A for limestone A, and Figure 6A for limestone B.

containing anhydrite. The same injection sequence as that


reported by Yousef et al.2 was used. Unlike limestone B, chalk
and limestone A show little IOR with the injection of diluted
seawater. Figure 7B shows the predicted adsorbed carboxylic
acid at the euent point. All three cores have decreased
Figure 6. (A) Oil recovery for forced displacement from core oods >CaOH2(COO) at 1.0 PVI with seawater injected. With 2
(Yousef et al.2) and simulation. (B) Euent concentration of chemical and 10 diluted seawater injection, limestone B shows
species. The units for Na+, total SO42 and free SO42 concentration decreased >CaOH2(COO), indicating a more water-wet
are mol/kg water. The units for >CaOH2(COO) concentration are state. Chalk and limestone A have an increased concentration
mol/m2. of >CaOH2(COO), indicating a more oil-wet rock because
H DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

the sulfate concentration decreased as seawater was diluted Sweep is excellent in the 2-D models, however, so that ultimate
(even after anhydrite was dissolved). Limestone A and oil recovery quickly approaches the 1-D values. This is expected
limestone B show signicantly dierent behavior. The dierent since the mobility ratio is favorable in these oods (end-point
IOR for limestone A and limestone B results from the use of mobility ratios are 0.5). For both the 1-D and 2-D cases, the
NEDL for limestone A, while EDL is used for limestone B. injection of 100 diluted FW produces 5% OOIP more oil
Figure 7C compares the oil recovery of the three cores with recovery than injection of FW at 2.0 PVI. This increase is
FW and 100 diluted FW injected. Both limestone A and signicant and demonstrates that recovery can be enhanced
limestone B show increased oil recovery, although the extent of with a realistic volume of water being injected.
the increase is much smaller in limestone A. The dierences in Figure 8B shows the oil recovery of SW and 10 diluted SW
IOR might be caused by shifts in relative permeability curves, injection for the same 1-D core and 2-D ve-spot models. For
which should be measured for each core. In contrast, no IOR is both injection uids, the 2-D ve-spot patterns exhibits smaller
observed for chalk. The >CaOH2(COO) concentrations in oil recovery at 2.0 PVI than the 1-D core oods. For both 1-D
Figure 7D indicate that wettability alteration occurs for and 2-D cases, injection of 10 diluted SW leads to 20%
limestones but does not occur for chalk, because of the lack OOIP more oil recovery at 2.0 PVI than the injection of SW
of anhydrite dissolution. The dierent responses of the chalk alone.
and limestones to the same injection water indicate the Figure 9A shows the permeability eld for a heterogeneous
importance of anhydrite dissolution. 2-D reservoir. The permeability eld contains a high-
Extension to the Field Scale Injection. One concern in
the core ood using LSW is the large pore volumes that are
injected (for example, 30 PVIs in Yousef et al.2) to achieve a
signicant increase in oil recovery. To examine this, we
simulated a quarter of a ve-spot pattern using PennSim with
the reaction network and parameters that reproduced core
ood experiments (see Tables 1 and 2). Two homogeneous
models and one heterogeneous model were simulated. The
pattern size of 300 ft 300 ft for all three models was
discretized into uniform 20 20 grid blocks. No gravity is
included. The rst two homogeneous models used the input
parameters of core ood experiments of Austad et al.5 and
Figure 9. Simulations for 2-D heterogeneous ve-spot pattern: (A)
Yousef et al.,2 respectively. The heterogeneous model used the permeability eld and (B) oil recoveries for 2-D ve-spot patterns
same porosity, mineral composition, and end-point relative (obtained using the same parameters as those in Figure 6B).
permeability of Yousef et al.2 and the permeability eld was
taken from an upscaling of the 41th layer of the permeability
eld in a SPE10 comparative solution project.57 The original
permeability was for 60 220 grid blocks and was coarsened permeability channel and therefore injection water break-
into 20 20 grid blocks. The grid coarsening was done by throughs occur much faster. Figure 9B shows a signicantly
combining 3 11 ne grid blocks into one coarse grid block. decreased oil recovery and earlier breakthrough, compared to
The permeability was calculated by using an arithmetic mean. the homogeneous case, because of poor sweep eciency,
PennSim results for the 2-D ve-spot pattern cases matched although the benet of using LSW remains large (6% OOIP).
exactly with simulations using Eclipse for FW injection without The CPU time for the entire simulation on a MacBook Pro
wettability alteration. (Intel i5 2.3 GHz, 2011 version) was 307 s.
Figure 8A compares the oil recovery curves for the 1-D Comparison with Previous Model. The reaction network
homogeneous core oods and 2-D homogeneous ve-spot used in this paper is extended from a previous model.8 With
patterns. Breakthrough in the 1-D simulations is delayed more reactions, the current model is compared with the
slightly compared to the 2-D model for both FW and 100 previous model for the case of ow in chalk and a series of
diluted FW injection water. Breakthrough is somewhat quicker injection brine compositions including formation water (FW),
in the 2-D homogeneous model, because of poorer sweep. seawater (SW), SW depleted in NaCl (SW0NaCl), and
SW0NaCl with four levels of sulfate (SW0NaCl4SO4). The
brine composition is shown in Table 3. As is shown in Figure
10 for the current model with and without calcite dissolution,
the simulations reproduce the decrease of >CaOH2(COO)
concentration on the rock surface. This indicates that the
current model and the previous model predict the same trend
of wettability alteration with the tested brine compositions. The
eect of calcite dissolution is also tested, since calcite
dissolution was regarded as being important during seawater
injection processes.58 Figure 10 also indicates that although
calcite dissolution is important when FW is injected, it does not
aect the trend of wettability alteration.
Figure 8. Oil recoveries for 1-D core oods and 2-D ve-spot patterns Model Limitations and Potential Improvements. The
in a homogeneous limestone reservoir, using (A) parameters of core model developed here oers a promising pathway to move
ood experiments from Austad et al.5 and (B) parameters of core forward with regard to understanding the chemical mechanisms
oods from Yousef et al.2 for wettability alteration during low-salinity waterooding. The
I DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

euent concentration matched well with the experimental


measurements.
(ii) Forced displacement experiments for the SK chalk and
limestone cores were also simulated. The predicted oil
recoveries also matched the experimental measurements well.
The predicted euent concentrations were consistent with
experimental measurements.
(iii) Simulation of secondary recoveries for the 1-D core
ooding models and 2-D ve-spot models ranged from 5% to
20% OOIP at 2.0 PVI, using diluted formation water or diluted
seawater. More importantly, the 2-D models demonstrated that
LSW can yield excellent increases in recovery at realistic pore
volumes injected (PVI).
Additional important conclusions include the following:
The same set of reactions can be used for the wettability
Figure 10. Comparison of the calculated surface fraction of alteration model.
>CaOH2(COO) for chalk from extended and previous models.46 The prediction of wettability alteration using the
Blue bars represent the values calculated by the previous model; red mechanistic model is accurate for the demonstrated
bars represent the values calculated by the current model; and green experimental studies.
bars represent the values calculated by the current model, but without Coreooding experiments are lacking all data needed to
calcite dissolution. predict recovery. Relative permeability curves should be
measured at two wettability states.
The nonelectrical double layer (NEDL) model cannot
current model does have some limitations that may aect the capture wettability alteration in the core oods of Yousef
predictive capability of this model. et al.2 The electrical double layer (EDL) model was
Four mechanisms of interaction between crude oil, brine, and required instead.
solid surfaces have been reported in the literature, including In limestone cores with anhydrite,5 diluted water reduces
polar, surface precipitation, acid/base and divalent/multivalent ion strength and aqueous complexation of sulfate with
ion binding.23 Here, we focus on the acid/base mechanism. We cations. This enhances anhydrite dissolution and the
did not include ion binding because there is a lack of availability of free SO42, thereby improving the oil
experimental and thermodynamic data for ion-bridging recovery. The simulation for the same limestone core but
interaction. Moreover, it is challenging to study those without anhydrite does not show IOR.
mechanisms independently. The IOR at 2.0 PVI for 1-D core oods and 2D ve-spot
The use of EDL and NEDL models to reproduce the models is dependent on the rock mineralogy.
experiments for limestone A5 and limestone B2 is not desirable.
This is the rst mechanistic model for carbonate rocks that
This indicates that the key dierence between dierent forms of
considers the coupling of mineral dissolution and surface
CaCO3, limestone and chalk, should be identied. Collection of
complexation reactions to predict low-salinity waterooding
detailed aqueous geochemical data, including Ca2+, Mg2+, and
recoveries and wettability changes for both chalk and limestone
SO42 concentrations, and pH at the euent, as well as
cores, with and without anhydrite. Other minerals could be
measuring zeta potential at the oil and rock surface at high
added as needed in the reaction network for reservoir
pressure and high temperature, will also allow further
formations with dierent type of lithology.


constraints on the geochemical reaction network.

CONCLUSIONS
In this research, we extended a previous mechanistic wettability
AUTHOR INFORMATION
Corresponding Author
*E-mail: changheqiao@gmail.com.
alteration model to understand and predict the improved oil
recovery (IOR) from low-salinity waterooding for dierent Notes
The authors declare no competing nancial interest.


cases, including seawater injection in the Stevns Klint chalk, and
diluted formation water and seawater injection in limestone
cores containing anhydrite. The extension of the reactions ACKNOWLEDGMENTS
include adding calcite and anhydrite dissolution, the sorption of The authors gratefully thank the member companies of the Gas
Ca2+ and Mg2+ on the calcite surface, and aqueous complex- Flooding JIP in the EMS Energy Institute at The Pennsylvania
ation. The new model was incorporated into an in-house State University at University Park, PA for their nancial
simulator PennSim, providing a unifying explanation of the support. Russell T. Johns holds the Victor and Anna Mae
recoveries observed under low-salinity waterooding conditions Beghini Faculty Fellowship in Petroleum and Natural Gas
Engineering at The Pennsylvania State University.


in carbonate reservoirs.
The results show that, with little tuning, the new model can
capture the role of mineral reactions in buering the aqueous NOMENCLATURE
composition and could accurately predict wettability alteration ai = activity of species i (dimensionless)
and recoveries, because of surface reactions. The main overall A, B = temperature-dependent parameters for the Debye
conclusions are as follows: Huckel model (kg water1/2/mol1/2)
(i) The 1-D chromatographic wettability tests for the SK d = temperature-dependent parameter for the Debye
chalk and limestone cores were simulated. The predicted Huckel model (kg water/mol)
J DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

As = bulk surface area of mineral species s (m2) (6) Fathi, S. J.; Austad, T.; Strand, S. Smart Water as a Wettability
D = diusion/dispersion coecient tensor (ft2/day or m2/s) Modifier in Chalk: The Effect of Salinity and Ionic Composition.
Ea = activation energy (Btu/lb-mol or J/mol) Energy Fuels 2010, 24 (4), 25142519.
F = Faradays constant; F = 9.648 104 C/mol (7) Zhang, P.; Tweheyo, M. T.; Austad, T. Wettability Alteration and
Improved Oil Recovery in Chalk: The Effect of Calcium in the
Fp = molar rate of the primary species p (mol/(m day))
Presence of Sulfate. Energy Fuels 2006, 20 (5), 20562062.
Fq = molar rate of the secondary species q (mol/(m day)) (8) Qiao, C.; Li, L.; Johns, R. T.; Xu, J. A Mechanistic Model for
g = standard gravitational constant (ft/day2 or m/s2) Wettability Alteration by Chemically Tuned Waterflooding in
I = ionic strength of water (mol/kg water) Carbonate Reservoirs. SPE J. 2015, 20 (4), 767783.
Keq,r = equilibrium constant of reaction r (9) Zhang, P.; Tweheyo, M. T.; Austad, T. Wettability Alteration and
K = permeability (md or m2) Improved Oil Recovery by Spontaneous Imbibition of Seawater into
ks = reaction rate constant (mol/m2s) Chalk: Impact of the Potential Determining Ions Ca2+, Mg2+, and
krj = relative permeability of phase j SO42. Colloids Surf., A 2007, 301 (1), 199208.
Mp = molar density of the primary species p (mol/m3) (10) Nasralla, R. A.; Sergienko, E.; Masalmeh, S. K.; van der Linde,
Mq = molar density of the secondary species q (mol/m3) H. A.; Brussee, N. J.; Mahani, H.; Suijkerbuijk, B.; Alqarshubi, I.
Np = number of primary species Demonstrating the Potential of Low-salinity Waterood to Improve
Oil Recovery in Carbonate Reservoirs by Qualitative Coreood. In
Ntot = total number of species
Abu Dhabi International Petroleum Exhibition and Conference, Abu
Nsec = number of secondary reactions Dhabi, UAE, Nov. 1013, 2014; Society of Petroleum Engineers:
Pj = pressure of phase j (psi or Pa) Richardson, TX, 2014.
Pcj = capillary pressure between phase j and a reference phase (11) Yousef, A. A.; Liu, J.; Blanchard, G.; Al-Saleh, S.; Al-Zahrani, T.;
(psi or Pa) Al-Zahrani, R.; Al-Tammar, H.; Al-Mulhim, N. Smartwater Flooding:
Qp = total molar rate of primary species p (lb-mol/day or Industrys First Field Test in Carbonate Reservoirs. In Proceedings of
mol/s) the 2012 SPE Annual Technical Conference and Exhibition; Society of
R = universal gas constant; R = 8.31 J/(mol K) or 10.73 ft3 Petroleum Engineers: Richardson, TX, 2012; pp 810.
psi/(R lb-mol) (12) Fathi, S. J.; Austad, T.; Strand, S. Water-Based Enhanced Oil
Sj = saturation of phase j (dimensionless) Recovery (EOR) by Smart Water: Optimal Ionic Composition for
T = temperature (R or K) EOR in Carbonates. Energy Fuels 2011, 25 (11), 51735179.
(13) Webb, K. J.; Black, C. J. J.; Tjetland, G. A Laboratory Study
Z = depth (ft or m)
Investigating Methods for Improving Oil Recovery in Carbonates.
zi = charge carried by species i Presented at the 2005 International Petroleum Technology Conference,
Greek Symbols Doha, Qatar, Nov. 2123, 2005; Conference Paper No. 10506-MS
0 = dielectric constant of water; 0 = 55.3 (dimensionless) IPTC.
m = permittivity of free space; m = 8.854 1013 C V1 (14) Hiorth, A.; Cathles, L.; Madland, M. The Impact of Pore Water
dm1 Chemistry on Carbonate Surface Charge and Oil Wettability. Transp.
Porous Media 2010, 85 (1), 121.
o = charge density at the oil surface (C/m2) (15) Tang, G.-Q.; Morrow, N. R. Influence of brine composition and
o = oil/water interface charge potential (mV) fines migration on crude oil/brine/rock interactions and oil recovery. J.
s = solid/water interface potential (mV) Pet. Sci. Eng. 1999, 24 (2), 99111.
j = molar density of phase j (lb-mol/ft3 or mol/m3) (16) Gomari, K. R.; Hamouda, A.; Denoyel, R. Influence of Sulfate
i = activity coecient of species i (kg water/mol) Ions on the Interaction Between Fatty Acids and Calcite Surface.
ip = the (i, p) entry in the stoichiometry matrix for the Colloids Surf., A 2006, 287 (1), 2935.
reactions in canonical form (17) Gupta, R.; Smith, G. G.; Hu, L.; Willingham, T.; Lo Cascio, M.;
= porosity (dimensionless) Shyeh, J. J.; Harris, C. R. Enhanced Waterood for Carbonate
= density (lb/ft3 or kg/m3) ReservoirsImpact of Injection Water Composition. In SPE Middle
j = viscosity of phase j (cP) East Oil and Gas Show and Conference, Manama, Bahrain, Sept. 2528,

2011; Society of Petroleum Engineers: Richardson, TX, 2011.


(18) Romanuka, J.; Hofman, J.; Ligthelm, D. J.; Suijkerbuijk, B.;
REFERENCES Marcelis, F.; Oedai, S.; Brussee, N.; van der Linde, H.; Aksulu, H.;
(1) Austad, T.; Strand, S.; Madland, M. V.; Puntervold, T.; Korsnes, Austad, T., Low Salinity EOR in Carbonates. In SPE Improved Oil
R. I. Seawater in Chalk: An EOR and Compaction Fluid. SPE J. 2008, Recovery Symposium, Tulsa, OK, USA, Apr. 1418, 2012; Society of
Petroleum Engineers: Richardson, TX, 2012.
11 (4), 648 (DOI: 10.2118/118431-PA).
(19) Chandrasekhar, S.; Mohanty, K. K. Wettability Alteration with
(2) Yousef, A.; Al-Saleh, S.; Al-Kaabi, A.; Al-Jawfi, M. S. Laboratory
Brine Composition in High Temperature Carbonate Reservoirs. In
Investigation of the Impact of Injection-Water Salinity and Ionic 2013 SPE Annual Technical Conference and Exhibition, New Orleans,
Content on Oil Recovery from Carbonate Reservoirs. SPE Reservoir LA, USA, Sept. 30Oct. 2, 2013; Society of Petroleum Engineers:
Eval. Eng. 2011, 14 (5), 578. Richardson, TX, 2013.
(3) Lager, A.; Webb, K. J.; Black, C. J. J.; Singleton, M.; Sorbie, K. S. (20) Jerauld, G. R.; Webb, K. J.; Lin, C.-Y.; Seccombe, J. C. Modeling
Low Salinity Oil RecoveryAn Experimental Investigation 1. Low-Salinity Waterflooding. SPE Reservoir Eval. Eng. 2008, 11 (6),
Petrophysics 2008, 49 (1), Paper No. SPWLA-2008-v49n1a2. 10001012.
(4) Yousef, A.; Al-Saleh, S.; Al-Kaabi, A.; Al-Jaw, M. S. Laboratory (21) Delshad, M.; Najafabadi, N. F.; Anderson, G.; Pope, G. A.;
Investigation of the Impact of InjectionWater Salinity and Ionic Sepehrnoori, K. Modeling Wettability Alteration by Surfactants in
Naturally Fractured Reservoirs. SPE Reservoir Eval. Eng. 2009, 12 (3),
Content on Oil Recovery From Carbonate Reservoirs. SPE Reservoir
361370.
Eval. Eng. 2011, 14 (5), 578 (DOI: 10.2118/137634-PA). (22) Al-Shalabi, E. W.; Sepehrnoori, K.; Delshad, M.; Pope, G. Novel
(5) Austad, T.; Shariatpanahi, S. F.; Strand, S.; Black, C. J. J.; Webb, A Method to Model Low-Salinity-Water Injection in Carbonate Oil
K. J. Conditions for a Low-Salinity Enhanced Oil Recovery (EOR) Reservoirs. SPE J. 2015, 20 (5), 11541166.
Effect in Carbonate Oil Reservoirs. Energy Fuels 2012, 26 (1), 569 (23) Buckley, J.; Liu, Y. Some Mechanisms of Crude Oil/Brine/Solid
575. Interactions. J. Pet. Sci. Eng. 1998, 20 (3), 155160.

K DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX
Energy & Fuels Article

(24) Qiao, C.; Li, L.; Johns, R. T.; Xu, J. Compositional Modeling of Interfaces and Their Spectroscopic Evaluation. Langmuir 2000, 16
Dissolution-Induced Injectivity Alteration During CO2 Flooding in (6), 26772688.
Carbonate Reservoirs. SPE J. 2015 (DOI: 10.2118/170930-PA). (45) Langmuir, D.; Hall, P.; Drever, J. Environmental Geochemistry;
(25) Yeh, G.-T.; Tripathi, V. S. A Model for Simulating Transport of Prentice Hall: Englewood Clis, NJ, 1997.
Reactive Multispecies Components: Model Development and (46) Qiao, C.; Johns, R.; Li, L.; Xu, J. Modeling Low Salinity
Demonstration. Water Resour. Res. 1991, 27 (12), 30753094. Waterooding in Mineralogically Dierent Carbonates. In 2015 SPE
(26) Steefel, C. I.; Lasaga, A. C. A Coupled Model for Transport of Annual Technical Conference and Exhibition, Houston, TX, USA, Sept.
Multiple Chemical Species and Kinetic Precipitation/Dissolution 2830, 2015; Society of Petroleum Engineers: Richardson, TX, 2015.
Reactions with Application to Reactive Flow in Single Phase (47) Brady, P. V.; Krumhansl, J. L. A Surface Complexation Model of
Hydrothermal Systems. Am. J. Sci. 1994, 294 (5), 529592. OilBrineSandstone Interfaces at 100 C: Low Salinity Water-
(27) Walter, A.; Frind, E.; Blowes, D.; Ptacek, C.; Molson, J. flooding. J. Pet. Sci. Eng. 2012, 81, 171176.
Modeling of Multicomponent Reactive Transport in Groundwater: 1. (48) Brady, P. V.; Morrow, N. R.; Fogden, A.; Deniz, V.; Loahardjo,
Model Development and Evaluation. Water Resour. Res. 1994, 30 (11), N. Electrostatics and the Low Salinity Effect in Sandstone Reservoirs.
31373148. Energy Fuels 2015, 29 (2), 666677.
(28) Steefel, C. I.; DePaolo, D. J.; Lichtner, P. C. Reactive transport (49) Hopkins, P.; Puntervold, T.; Strand, S. Preserving Initial Core
modeling: An essential tool and a new research approach for the Earth Wettability During Core Restoration of Carbonate Cores. Presented at
Sciences. Earth Planet. Sci. Lett. 2005, 240, 539558. the 29th International Symposium of the Society of Core Analysts, St.
(29) MacQuarrie, K. T. B.; Mayer, K. U. Reactive transport modeling Johns Newfoundland and Labrador, Canada, 2015.
in fractured rock: A state-of-the-science review. Earth-Sci. Rev. 2005, (50) Strand, S.; Puntervold, T.; Austad, T. Effect of Temperature on
72 (34), 189227. Enhanced Oil Recovery From Mixed-wet Chalk Cores by Spontaneous
(30) Bao, C.; Wu, H.; Li, L.; Newcomer, D.; Long, P. E.; Williams, K. Imbibition and Forced Displacement Using Seawater. Energy Fuels
H. Uranium Bioreduction Rates across Scales: Biogeochemical Hot 2008, 22 (5), 32223225.
Moments and Hot Spots during a Biostimulation Experiment at Rifle, (51) Zahid, A.; Shapiro, A. A.; Skauge, A. Experimental Studies of
Colorado. Environ. Sci. Technol. 2014, 48 (17), 1011610127. Low Salinity Water Flooding Carbonate: A New Promising Approach.
(31) Brunet, J. P. L.; Li, L.; Karpyn, Z. T.; Kutchko, B. G.; Strazisar, In 2012 SPE EOR Conference at Oil and Gas, West Asia, 2012; Society
B.; Bromhal, G. Dynamic Evolution of Cement Composition and of Petroleum Engineers: Richardson, TX, 2012.
Transport Properties under Conditions Relevant to Geological Carbon (52) Mahani, H.; Keya, A. L.; Berg, S.; Bartels, W.-B.; Nasralla, R.;
Sequestration. Energy Fuels 2013, 27 (8), 42084220. Rossen, W. R. Insights into the Mechanism of Wettability Alteration
(32) Surasani, V. K.; Li, L.; Ajo-Franklin, J. B.; Hubbard, C.; by Low-Salinity Flooding (LSF) in Carbonates. Energy Fuels 2015, 29
Hubbard, S. S.; Wu, Y. Bioclogging and Permeability Alteration by L. (3), 13521367.
mesenteroides in a Sandstone Reservoir: A Reactive Transport (53) Brooks, R. H.; Corey, A. Properties of Porous Media Affecting
Modeling Study. Energy Fuels 2013, 27 (11), 65386551. Fluid Flow. J. Irrig. Drain. Div., Am. Soc. Civ. Eng. 1966, 92 (2), 6190.
(33) Vilcaez, J.; Li, L.; Wu, D.; Hubbard, S. S. Reactive Transport (54) Morrow, N. R.; Lim, H. T.; Ward, J. S. Effect of crude-oil-
Modeling of Induced Selective Plugging by Leuconostoc mesenteroides in induced wettability changes on oil recovery. SPE Form. Eval. 1986, 1
Carbonate Formations. Geomicrobiol. J. 2013, 30 (9), 813828. (1), 89103.
(34) Hubbard, C. G.; Cheng, Y.; Engelbrekston, A.; Druhan, J. L.; Li, (55) Sorop, T. G.; Masalmeh, S. K.; Suijkerbuijk, B. M. J. M.; van der
L.; Ajo-Franklin, J. B.; Coates, J. D.; Conrad, M. E. Isotopic insights Linde, H. A.; Mahani, H.; Brussee, N. J.; Marcelis, F. A. H. M.; Coorn,
into microbial sulfur cycling in oil reservoirs. Front. Microbiol. 2014, 5 A. Relative Permeability Measurements to Quantify the Low Salinity
(DOI: 10.3389/fmicb.2014.00480). Flooding Eect at Field Scale. Presented at the Abu Dhabi International
(35) Morel, F.; Hering, J. G. Principles and Applications of Aquatic Petroleum Exhibition and Conference, Abu Dhabi, UAE, Nov. 912,
Chemistry; Wiley: New York, 1993; pp xv, 588. 2015; Paper No. SPE-177856-MS (DOI: 10.2118/177856-MS).
(36) Lasaga, A. C. Chemical kinetics of water-rock interactions. J. (56) Strand, S.; Austad, T.; Puntervold, T. Smart Water for Oil
Geophys. Res. 1984, 89 (B6), 40094025. Recovery from Fractured Limestone: A Prelimlinary Study. Energy
(37) Wolery, T. J. EQ3/6: A Software Package for Geochemical Fuels 2008, 22, 3126.
Modeling of Aqueous Systems: Package Overview and Installation Guide (57) Christie, M.; Blunt, M. Tenth SPE Comparative Solution
(Version 7.0); Lawrence Livermore National Laboratory: Livermore, Project: A Comparison of Upscaling Techniques. In 2001 SPE
CA, 1992. Reservoir Simulation Symposium, Houston, TX, Feb. 1114, 2001;
(38) Helgeson, H. C.; Brown, T. H.; Nigrini, A.; Jones, T. A. Society of Petroleum Engineers: Richardson, TX, 2001.
Calculation of Mass Transfer in Geochemical Processes Involving (58) Andersen, P.; Evje, S.; Madland, M.; Hiorth, A. A geochemical
Aqueous Solutions. Geochim. Cosmochim. Acta 1970, 34 (5), 569592. model for interpretation of chalk core flooding experiments. Chem.
(39) Lasaga, A. C. Kinetic Theory in the Earth Sciences; Princeton Eng. Sci. 2012, 84, 218241.
(59) Chou, L.; Garrels, R. M.; Wollast, R. Comparative study of the
University Press: Princeton, NJ, 2014.
(40) Brady, P. V.; Krumhansl, J. L.; Mariner, P. E. Surface kinetics and mechanisms of dissolution of carbonate minerals. Chem.
Geol. 1989, 78, 269282.
Complexation Modeling for Improved Oil Recovery. In 2012 SPE
(60) Jeschke, A. A.; Vosbeck, K.; Dreybrodt, W. Surface controlled
Improved Oil Recovery Symposium, Tulsa, OK, USA, Apr. 1418, 2012;
dissolution rates of gypsum in aqueous solutions exhibit nonlinear
Society of Petroleum Engineers: Richardson, TX, 2012.
dissolution kinetics. Geochim. Cosmochim. Acta 2001, 65 (1), 2734.
(41) Qiao, C.; Li, L.; Johns, R. T.; Xu, J., A Mechanistic Model for
(61) Shariatpanahi, S. F.; Strand, S.; Austad, T. Evaluation of Water-
Wettability Alteration by Chemically Tuned Water Flooding in
Based Enhanced Oil Recovery (EOR) by Wettability Alteration in a
Carbonate Reservoirs. In 2014 SPE Annual Technical Conference and
Low-Permeable Fractured Limestone Oil Reservoir. Energy Fuels 2010,
Exhibition, Amsterdam, Oct. 2729, 2014; Society of Petroleum
24 (11), 59976008.
Engineers: Richardson, TX, 2014.
(42) Pokrovsky, O.; Schott, J. Surface Chemistry and Dissolution
Kinetics of Divalent Metal Carbonates. Environ. Sci. Technol. 2002, 36
(3), 426432.
(43) Goldberg, S. Surface complexation modeling. Reference Module
Earth Syst. Environ. Sci. 2013, 10 (DOI: 10.1016/B978-0-12-409548-
9.05311-2).
(44) Pokrovsky, O. S.; Mielczarski, J. A.; Barres, O.; Schott, J. Surface
Speciation Models of Calcite and Dolomite/Aqueous Solution

L DOI: 10.1021/acs.energyfuels.5b02456
Energy Fuels XXXX, XXX, XXXXXX

Anda mungkin juga menyukai