An Experimental Investigation Into The Impact of Sulfate Ions in Smart Water To Improve Oil Recovery in Carbonate Reservoirs

You might also like

You are on page 1of 20

Transp Porous Med

DOI 10.1007/s11242-015-0616-4

An Experimental Investigation into the Impact of Sulfate


Ions in Smart Water to Improve Oil Recovery in
Carbonate Reservoirs

Adedapo Awolayo1 · Hemanta Sarma1 · Ali AlSumaiti2

Received: 4 April 2015 / Accepted: 7 December 2015


© Springer Science+Business Media Dordrecht 2015

Abstract Carbonate reservoirs are considered to be in the range of mixed to oil wetting
state. Their present wetting state seems unfavorable to recover more oil via conventional
waterflooding especially as the water cut increases with huge residual oil saturation. In recent
years, studies have shown that tuning the injected brine chemistry can help decrease the oil-
wetness, resulting in additional oil recovery and reduction in residual oil saturation. Thus,
rock wettability is being dictated by the surface chemistry associated with the water-film
stability between crude oil and the rock surface. This study presents series of experiments
conducted on Middle Eastern carbonate core plugs at high pressure and high temperature in
order to determine the impact of sulfate ions in smart brine on oil recovery. Each experiment
was initially conducted using the formation brine, and subsequently, various versions of
seawater with varying sulfate concentrations were used. Coreflooding experiments showed
an additional recovery of about 7–9 % oil originally in core when sulfate concentration was
increased while none when decreased. Decrease in rock oil-wetness was observed to be
responsible for the improved recovery as demonstrated by the contact angle analysis. Similar
trend was observed with the zeta potential analysis which further proved that the magnitude
of surface charge alteration contributes immensely to wettability modification toward more
water-wetness.

Keywords Smart water · Wettability · Zeta potential · Improved recovery · Carbonate

List of Symbols
D Diameter
IRF Incremental recovery factor
IS Ionic strength

B Adedapo Awolayo
adedapoawolayo@yahoo.com
1 Department of Chemical and Petroleum Engineering, University of Calgary,
2500 University Dr. NW, Calgary, AB T2N 1N4, Canada
2 Petroleum Engineering Department, The Petroleum Institute, P.O. Box 2533, Abu Dhabi, UAE

123
A. Awolayo et al.

Kg Air permeability
KL Liquid permeability
K o (Swirr ) Effective permeability to oil at irreducible water saturation
kppm Thousands part per million
L Length
∅ Porosity
OOIC Original oil in core
PV Pore volume
RF Recovery factor
Sor Residual oil saturation
Swirr Irreducible water saturation
TDS Total dissolved solids
ζ Zeta

1 Introduction

Carbonate reservoirs were reported to contain almost half of the world hydrocarbon proven
reserves with 90 % of them observed to be mixed-wet to oil-wet (Sheng 2013a). Relating few
features like reservoir management, fluid types, reservoir heterogeneities and drive mech-
anisms, experts presumed that more than 50 % of oil trapped in carbonate rocks are yet to
be recovered (Rod 2003). Thus far, different techniques of enhanced oil recovery (EOR)
have been applied to carbonate formations to reduce residual oil saturation (Sor ) and improve
recovery.
Over the years, waterflood had been one of the dominant and successful fluid injection
methods for oil recovery and this has been attributed to water availability. However, water
injection is relatively cheap and efficient in displacing light/medium API oil. The function of
the injected water chemistry as well as its influence on oil recovery has received less attention
because it is seen as a physical method of maintaining pressure and oil displacement. Smart
waterflooding is a novel technique that can improve the waterflood performance through the
addition and/or removal of ions from the injection water. This concept relies on the belief that
both the physical and chemical properties of the water are of great significance for improving
oil recovery.
Extensive research, on both sandstones (Jadhunandan and Morrow 1995; Lager et al. 2007,
2008; Tang and Morrow 1999) and carbonates (Austad 2013; Austad et al. 2011; Strand et al.
2006), has shown promising outcome of modifying the composition and/or salinity of the
injected brine on crude oil/brine/rock (CBR) interactions by wettability alteration process.
The observed alteration augments the microscopic displacement efficiency and therefore
improves oil recovery.
Successful field trials (McGuire et al. 2005; Webb et al. 2004) have shown that smart
waterflooding can improve the oil recovery in sandstone reservoirs. The reported mechanism
observed for the improved recovery by smart water is the modification of wettability of
the clay minerals present in sandstones which is considered a key prerequisite for smart
waterflooding application (Ligthelm et al. 2009).

1.1 Smart Water Application in Carbonate

Experimental feat reported on carbonates, following a paradigm shift to carbonate reservoirs,


had weakened the argument that smart water would not improve oil recovery in carbonate due

123
An Experimental Investigation into the Impact of Sulfate. . .

to the absence of clay minerals (Lager et al. 2008). RezaeiDoust et al. (2009) had reported that
the success achieved was as a result of contrasting chemical mechanisms between carbon-
ates and sandstones. These mechanisms involved crude oil attachment to the calcite surface
(charged positive) and quartz surface (charged negative), respectively. Moreover, the bond-
ing energy existing between polar components of the crude oil and the rock surface is lesser
in sandstones compared to carbonates. Besides, it was reported that brine (salinity up to
33 kppm) altered wettability in carbonates, while brine (very low salinity in the range of
2000–3000 ppm) did same in sandstones. This fact was supported by the occurrence of ionic
exchange between the rock and the high salinity invading brine. This process only occurred
when active ions concentrations are varied between the formation brine and invading brine.
After years of research, Austad and his coworkers identified the type and relative con-
centration of the active ions as one of the elements responsible for wettability alteration
(Strand et al. 2006; Zhang et al. 2007). Zhang and Austad (2006) identified SO2− 4 , Ca
2+ and
2+
Mg as the active ions in seawater that altered the rock surface charge and increased the
water-wetness of the rock. Høgnesen et al. (2005) conducted a spontaneous imbibition exper-
iment to check the effect of sulfate on the alteration of preferentially oil-wet chalk toward
more water-wetness. Results obtained showed increase in oil recovery with increased sulfate
concentration. Sulfate, which is believed to be a very strong active ion, acts as a wettability
modifier toward carbonate surface (Pierre et al. 1990).
It was proposed that sulfate adsorption onto the chalk surface lowered the positive charge
and created an increased electrostatic attraction between the chalk surface and negatively
charged oil. This attraction allowed the localization of excess Ca2+ closer to the surface,
which reacted and desorbed the oil polar component from the chalk surface. This process
enhanced the change in the rock wettability toward more water-wetness. Zhang et al. (2006)
observed that the symbiotic effect of Ca2+ and SO2− 4 on the wettability modification process
is temperature-dependent. As the temperature increased, these active ions developed more
reactivity to the rock surface, and this compelled Mg2+ to displace Ca2+ in the calcium
carboxylate complex (Zhang et al. 2007). Because of dehydration at higher temperatures,
Mg2+ developed more reactivity and displaced Ca2+ at the rock surface. Similar to chalk
reservoirs, the same trend was observed in limestone reservoirs (Strand et al. 2008). In
another set of experiments on carbonate rocks, Gupta et al. (2011) replaced sulfate ions in
seawater with borate and phosphate ions. They observed additional recoveries of 21 and
16 %, respectively.
Another approach suggested besides increasing active ions concentrations in the injec-
tion brine was ionic strength reduction and as well demonstrated rock wettability alteration.
Yousef et al. (2010) reported 16–18 % oil originally in core (OOIC) increment while inves-
tigating the impact of salinity reduction during smart waterflooding. The experiments were
performed at reservoir conditions and materials used were: Saudi Arabian carbonate core
plugs placed in series (composite), live oil and Gulf seawater. The identified driving mecha-
nism for incremental recovery was wettability alteration due to rock surface charge alteration.
Austad et al. (2011) performed a study to investigate low-salinity effects on carbonate
reservoirs reported by Yousef et al. (2010). They observed an incremental 2–5 % OOIC as
the brine ionic strength reduced and anhydrite was flushed out from the sample at the end
of the experiment. They later concluded that reduction of ionic strength can improve rock
water-wetness and oil recovery on core samples containing anhydrite. Zahid et al. (2012)
had conducted a similar experiment but observed no incremental recovery below 90 ◦ C and
proposed fines migration to be the possible mechanisms for oil recovery. Previous study
conducted by Yi and Sarma (2012) also showed an incremental recovery when the salinity
of seawater was reduced and with increased sulfate concentration.

123
A. Awolayo et al.

As articulated in the literature, the wetting state of carbonate rocks was altered by two
approaches supported by experimental evidences under certain conditions:

1. Increase in the surface interacting ion concentration, i.e., phosphate, borate, sulfate in
the brine,
2. Reduction in the salinity/ionic strength of the brine.
(a) Reduction of cation concentration,
(b) Removal of non-active ions, i.e., Na+ , Cl− from the brine,
(c) Brine dilution.

Based on the above discussion, this present study was built to investigate the potential of
varying sulfate concentration in smart water on oil recovery in carbonate reservoirs, par-
ticularly limestone. The fundamental objective was to develop a rigid understanding of the
impact of sulfate ions during smart waterflooding in carbonate and deduce how this ion alters
rock wettability.

2 Experimental Studies

2.1 Materials

2.1.1 Fluids

All synthetic brines were prepared in the laboratory by dissolving specified amounts of
reagent-grade salts in deionized water. The geochemical formulations of different brines
used are given in Table 1. The brines used in this paper are classified as either smart brine
or base brine. Smart brines were injected after base brine and possibly to trigger incremental
recovery. Smart brines were prepared based on seawater compositions by maintaining the
ionic strength, and their nomenclature reflects the relative change in the active ions concen-
trations. Therefore, formation brine was tagged “FB,” synthetic seawater brine as “SSB,”
seawater with half of the typical SO2−
4 concentration as “SSB#0.5S,” synthetic seawater with
two times the typical SO2−4 concentration as “SSB#2S,” synthetic seawater with four times
the typical SO2−
4 concentration as “SSB#4S” and synthetic seawater with eight times the typ-
ical SO2−
4 concentration as “SSB#8S.” FB was used as the base brine to establish irreducible
water saturation and secondary flooding.

Table 1 Geochemical analysis of the synthetic brines

Brines/ions (kppm) Na+ Ca2+ Mg2+ K+ SO2−


4 HCO−
3 Cl− TDS IS

FB 76.68 19.12 3.35 0.08 0.11 0.06 161.81 261.21 5.19


SSB 13.64 0.39 1.73 0.01 2.85 0.01 24.65 43.28 0.87
SSB # 0.5S 14.70 0.52 1.62 0.01 1.66 0.00 24.87 43.37 0.87
SSB # 2S 12.21 0.55 1.87 0.01 6.69 0.01 20.32 41.66 0.88
SSB # 4S 11.99 0.64 2.02 0.02 14.22 0.01 15.01 43.90 0.97
SSB # 8S 12.97 0.49 2.06 0.01 27.09 0.01 6.88 49.51 1.14

123
An Experimental Investigation into the Impact of Sulfate. . .

Table 2 Brine fluid properties

FB SSB SSB#0.5S SSB#4S SSB#8S SSB#2S

Density (g/cm3 ) @ 20 ◦ C 1.1728 1.0306 1.0309 1.032 1.0409 1.0301


Viscosity (cP) @ 70 ◦ C 0.815 0.463 0.458 0.474 0.488 0.467

Table 3 Reservoir fluid properties of oil samples

Analysis Results SARA analysis Results

Wax content (% m/m) 6.90 Saturates fraction (wt%) 49.5


Pour point (◦ C) −3 Aromatics fraction (wt%) 40.2
Sulfur (% m/m) 0.9 Resins fraction (wt%) 7.2
Wax appearance temperature (◦ C) 22 Asphaltenes fraction (wt%) 3.1
Density (g/cm3 ) @ 20 ◦ C 0.8376 Hydrocarbons/non-hydrocarbons ratio (wt./wt.) 8.7
Viscosity (cP) @ 70 ◦ C 1.927

Table 4 Basic petrophysical properties of carbonate core plugs

Sample ID Length (cm) Diameter (cm) ∅ (%) K L (mD) Bulk volume (cm3 ) Pore volume (cm3 )

S#2 6.23 3.86 20.05 1.59 72.73 14.58


S#3 5.75 3.84 27.09 4.52 66.55 18.03
S#4 6.49 3.86 19.84 1.63 76.03 14.95
S#9 6.22 3.84 24.66 7.31 72.23 17.94
S#13 6.12 3.80 33.24 3.60 69.17 22.99

The properties of the dead oil and brine are listed in Tables 2 and 3. The brines and
reservoir oil were filtered through 5-µm filter paper to remove solid particles that might
cause plugging problem.
SARA analysis of the crude oil sample showed saturates and aromatics ratio of 1.2 wt%
as well as resins and asphaltene ratio of 2.3 wt%, which indicated the presence of high polar
components. The acid number was not measured, and therefore, the crude oil acidity could
not be quantified. However, as an indication of the oil acidity, the rock plate showed strong
oil-wet characteristic at the initial state and could only change to a slightly oil-wet state as
discussed in Sect. 3.2.

2.1.2 Core plugs

Cores with consistent petrophysical properties were selected based on routine core analysis.
This was first conducted to measure the dimensions, permeability (K L ) and porosity (∅) of
the core plugs. Table 4 shows detailed petrophysical properties. Based on the core analysis
review, two composite cores were selected for the experiment. Rock plates cut from plugs
into rectangular shape of dimension (1.3 cm × 1.9 cm × 0.8 cm) were used for the contact
angle measurements. X-ray diffraction (XRD) analysis was conducted to identify the rock
samples, and results showed calcite with little trace of minerals like dolomite and quartz.

123
A. Awolayo et al.

2.2 Apparatus and Procedures

2.2.1 Zeta Potential Experiment

Zeta electroacoustic spectrometer was applied to measure zeta (ζ ) potential for rock/brine
interface with measurement accuracy of ±0.1 mV + 0.5 %. The instrument has a probe elec-
trode coated with palladium. It determined the ζ -potential and particle size distribution using
an acoustophoresis method that presents an advantage of much higher concentration over
the electrophoresis method (Awolayo et al. 2014b). The measurements were conducted at
25 ◦ C. Carbonate rocks were pulverized to very fine particles and sieved in order to maintain a
uniform particle size (< 15 µm). A representative solution sample was replicated by mixing
the FB with 12 wt% pulverized particle overnight using the magnetic stirrer.
Cuvette was used to hold about 25 mL of each representative solution sample; while stirring
at constant rate, the probe was inserted to obtain the initial charge on the rock surface. The
average value of three runs of measurements was recorded. Since ζ -potential is pH-dependent,
in this work, pH was not in any way manipulated so the corresponding pH was measured
as well. Then, various smart brines were mixed with each representative sample and the
ζ -potential measurement was retaken. With documented studies (Takamura and Chow 1985)
showing that the oil/brine interface is strongly negative at pH > 3, it was therefore imperative
to obtain the charge at the rock/brine interface. This would help to understand the relationship
between surface charge variation and oil recovery.

2.2.2 Contact Angle Measurement

The rock plates were polished with sand paper in order to reduce the hysteresis effect on the
rock surface due to surface roughness that could significantly influence the measurements.
The polished rock plates were placed under vacuum for at least 2 h and aged in the FB for
another 24 h. The plates were later aged in the reservoir oil at reservoir temperature of 110 ◦ C
for at least 6 weeks to restore the plates’ wettability toward reservoir conditions. Because of
the limitation to water boiling at 100 ◦ C, the test was carried out at 95 ◦ C, so the brines were
preheated to 95 ◦ C before commencement.
After aging, the plates were subjected to FB in order to verify the plates’ wettability,
and after observing the contact angle to be oil-wet (i.e., oil stuck to the plate), the plate
was then immersed in various preheated brines including the base brine. An oil drop was
injected to the plate surface and kept in a sealed vessel in the oven. Confined drop of crude
oil was seen to be formed on the rock plates with time. The contact angle of the drop usually
display time-variant behavior, and the evolution of these angles were monitored for a specific
period. Digital photographs were taken at each time interval. The advanced contact angle was
only subjected to quantitative analysis by measuring the angle between the baseline and the
droplet’s tangent. The rock plate was considered water-wet for 0◦ < θ < 70◦ , neutral-wet for
70◦ < θ < 110◦ and oil-wet for 110◦ < θ < 180◦ (Anderson 1986), while weakly water-wet
and weakly oil-wet behavior were considered as 55◦ < θ < 75◦ and 115◦ < θ < 135◦ ,
respectively.

2.2.3 Coreflooding Experiment

The coreflood apparatus consisted of components like Hassler-type core-holder, differential


pressure transducers, overburden pressure module, backpressure regulator (BPR), displace-

123
An Experimental Investigation into the Impact of Sulfate. . .

Fig. 1 Coreflooding apparatus schematics (Awolayo et al. 2014b)

Table 5 Coreflood injection sequence

Sample PV (cm3 ) OOIC (cm3 ) Swirr (%) Secondary Tertiary Tertiary Tertiary

S#9 17.82 14.00 21.42 FB SSB SSB#0.5S


S#42 23.16 18.43 20.44 FB SSB SSB#4S
S#313 39.54 31.50 20.33 FB SSB#2S SSB#4S SSB#8S

ment pumps, oven and fluid accumulators as labeled in Fig. 1. The system is capable of
handling maximum temperature of 200 ◦ C: overburden and pore pressures up to 69 MPa.
Each core plug was cleaned using toluene and methanol by Soxhlet extraction, dried at
a temperature of 130 ◦ C for 24 h and weighed. After routine core analysis, each core plug
was saturated under vacuum for 3 days with FB to establish ionic equilibrium with the brine
and ensure full brine saturation. The wet weight of each core plug was measured, and the
pore volume was computed from the difference in weight before and after saturation. Each
saturated core was flooded with reservoir oil until water production ceased which indicated
the establishment of the irreducible water saturation (Swirr ).
The oil permeability at irreducible water saturation was also measured, and the core plugs
were aged in the reservoir oil at 110 ◦ C and 13.8 MPa for at least 6 weeks to restore rock
wettability. For the waterflood tests, the core sample was placed inside a rubber sleeve to a
central position, mounted in the core-holder and connected at both ends with two end plugs
(one to the inlet and another to the outlet flow-line, respectively). Reservoir oil and brines
were supplied from the piston accumulators operated by high-pressure displacement pumps.
An overburden pressure was applied on the core by injecting hydraulic oil into the annulus
between the rubber sleeve and inner surface of the core-holder. Pore pressure was maintained
by a backpressure regulator (BPR) and measured by the differential pressure transduc-
ers. Each waterflood test commenced with the base brine at a constant injection rate of
0.25 mL/min. Once the oil production ceased, it was followed by sequential injection of
smart brines based on the sequence listed in Table 5. All experiments were conducted at
110 ◦ C, 20.7 MPa pore pressure and 10.3 MPa net confining pressure with horizontal core
orientation.
The produced fluid was collected to quantify the oil recovery, and ionic analysis for active
ions (Ca2+ , Mg2+ , and SO2+ + −
4 ) and non-active ions (Na and Cl ) was conducted for the
effluent brine. Collected brine samples at specific brine injected pore volumes were analyzed
by ion chromatography. Ionic analysis comparison was done by normalizing the effluent

123
A. Awolayo et al.

Table 6 Core plugs arrangement for coreflooding

Sample L (cm) D (cm) ∅ (%) K g (mD) K L (mD) PV (cm3 ) OOIC (cm3 ) Swirr (%) K o (Swirr ) (mD)

Single core
S#9 6.22 3.84 24.66 9.11 7.31 17.82 14.00 21.42 2.050
Composite core
S#4 3.70 3.86 19.84 1.19 0.82 8.60 7.23 15.99
S#2 6.23 3.86 20.05 2.05 1.59 14.56 11.20 23.06
S#42 9.93 3.86 19.95 1.62 1.21 23.16 18.43 20.44 0.497
Composite core
S#3 5.75 3.84 27.09 5.74 4.52 17.43 13.50 22.54
S#13 6.12 3.80 33.24 4.56 3.60 22.11 18.00 18.59
S#313 11.87 3.82 30.17 5.15 4.06 39.54 31.50 20.33 0.724

brine concentration against the injected brine concentration at specific brine injected pore
volume. At the end of the coreflood, the core plugs were subjected to Dean–Stark extraction,
and a material balance calculation was done to cross-check the recorded results.

3 Results and Discussion

This section presents the critical discussion of the series of experiments conducted on cores
itemized in Table 6. In all experiments, FB was set as baseline to compare the performance
of various smart brines.

3.1 Zeta Potential

The charges at oil/brine and rock/brine interfaces are the prime element that controls the
water-film stability between the oil and rock, hence the rock wettability. Stable and thick
water film bounded by the two interfaces would result in water-wet rock, while unstable and
thin water film would change the rock wettability toward oil-wet (Sheng 2013b). The rock
wettability depends on the sign and magnitude of the electrical charges at the two interfaces
due to the electrostatic (attractive or repulsive) forces generated.
Solution pH was reported with the ζ -potential measurement in Fig. 2 as discussed in
Sect. 2.2. The result showed that the charges at the rock/brine interface were positive at
all range of salinity. For the baseline (salinity 261 kppm), ζ -potential measured a positive
value of 4.49 mV ± 0.15 at a solution pH of 7.33. Then, mixing SSB#0.5S (42 kppm) with
the suspension gave a positive ζ -potential value of 4.29 mV ± 0.06 at a solution pH of 7.52
while SSB gave 2.22 mV ± 0.04 at solution pH of 7.86. Introduction of SSB#2S (43 kppm)
gave 1.99 mV ± 0.07 at solution pH of 7.55. In addition, SSB#4S gave 1.72 mV ± 0.04 at a
solution pH of 7.58, while SSB#8S gave 2.1 mV ± 0.07 at a solution pH of 8.34.
Interesting to note is that as the sulfate concentration increased, the interface charge
changed to less positive except SSB#8S. It is widely known that increasing solution pH leads
to reduction in surface charges due to less Ca2+ and more CO2− 3 species that are present at
high pH (Farooq et al. 2011). Since SO2− 4 has higher electronegativity than CO2−
3 , it displaced
2−
CO3 from the interface, increased the solution pH and reduced the surface charge at the

123
An Experimental Investigation into the Impact of Sulfate. . .

Zeta pH
5 8.6

8.4
4 8.2
Zeta Potenal (mV)
8
3
7.8

pH
7.6
2
7.4

1 7.2

0 6.8
FB SSB#0.5S SSB SSB#2S SSB#4S SSB#8S
Brine

Fig. 2 ζ -Potential as a function of pH with increasing [SO2−


4 ]

rock/brine interface. However, SSB gave a contrary solution pH trend. The magnitude of
the surface charge reduction could not be compared to the increase in sulfate concentration
beyond SSB, though the little more reduction was significant.
Increasing sulfate concentration in smart brine was expected to lower solution pH because
excess H+ is created into the aqueous interface during water dissociation. Nevertheless, as
smart brines interacted with the rock, the carboxylic group of the oil detached from the
rock surface together with excess divalent cations. This created an excess Cl− and SO2− 4 at
the interface which combined with the free H+ to reduce the free H+ concentration, thus
increased the solution pH toward more alkalinity.
The obtained results trend in Fig. 2 could be explained as follows: The carbonate rock
carried positive charges, while FB containing high cations concentration carried positive
charges as well. This made the interface to be dominantly positively charged. With smart
brine of increased sulfate concentration, the negatively charged sulfate dictated the aqueous
solution at high pH and as a result lowered the positive charge on the rock surface (Awolayo
et al. 2014a; Kwak et al. 2014).
Reducing the charges at the rock/brine interface with a strong negatively charged oil/brine
interface created an increased repulsive force between the interfaces. The stronger the mag-
nitude of this charge, the greater the electrostatic repulsive forces. Therefore, this repulsion
generated the expansion of the electrical double layer, which stabilized and thickened the
water film surrounding the rock. As a result, the rock wettability became less oil-wet. Gen-
erally, the significant effect of increasing sulfate concentration in smart brine resulted in the
stability and thickness of the water film.
Meanwhile, the less efficient effect generated beyond four times SO2− 4 concentration was
undoubtedly due to the increased precipitation reactions that occurred as sulfate concentration
increased. The precipitation reduced the amount of sulfate accessible to compete with CO2− 3
for the rock surface. Therefore, it is too early to conclude that SSB#4S would be the optimum
smart brine in this case.

3.2 Wettability Monitoring

The effect of various smart brines on carbonate rock wettability was tested by measuring the
contact angle of oil droplet on the rock plates. The prime objective was to investigate the

123
A. Awolayo et al.

FB SSB#0.5S SSB SSB#2S SSB#4S SSB#8S


180

170

160
Contact angle (deg.)

150

140

130

120

110

100

90
0 20 40 60 80 100 120
Time (hours)

Fig. 3 Contact angle versus different time interval: varying [SO2−


4 ] in smart water

wettability alteration by smart water as the principal mechanism responsible for improved
oil recovery. In addition, this test was designed to further support the evidence of surface
charge alteration as identified through ζ -potential measurement. The contact angles plotted
against exposure time are shown in Fig. 3.
Six different experiments were performed at 95 ◦ C with the same procedure as mentioned
in Sect. 2.2. For the initial time (as shown in Fig. 4), the rock plate showed oil-wet charac-
teristics at first contact with various brines after aging. The oil spread conveniently on the
plate surface and the surrounding brine could not directly contact the surface. At this stage,
it was deduced that the electrostatic repulsive forces between the two interfaces could not
be sustained by FB. Therefore, the thick water film became unstable, thinner and eventu-
ally ruptured, which made oil to contact the plate surface and change the wetting state into
strongly oil-wet.
Figure 4a shows the wettability-monitoring test performed under long exposure to same
high-saline brine. FB drastically decreased the contact angle to 151.2◦ during the first 15 h
and finally decreased to 135◦ , which was still in the range of preferentially oil-wet state.
The high salinity and reduced concentration of active ions in FB were not able to weaken
the strong electrostatic attractive forces. Thus, the water film was not thickened and stable
enough to detach the oil from the rock surface. This rendered the rock surface preferentially
oil-wet, which was responsible for high residual oil saturation.
As the exposure time to brine (with increased SO2− 4 concentration) increased, the contact
angle decreased; SSB#0.5S (Fig. 4b) reached a final value of 127.6◦ ; SSB (Fig. 4c) reached
a final value of 119.8◦ ; SSB#2S (Fig. 4d) reached a final value of 121.7◦ ; SSB#4S (Fig. 4e)
reached a final value of 100.3◦ ; SSB#8S (Fig. 4f) reached a final value of 121.7◦ . The plate
surface was changed from strongly oil-wet to weakly oil-wet state. This meant that the water
film on the rock surface became stable and thicker and hence prevented oil contact with the
plate surface.
The result indicated that as the SO2−
4 concentration increased, an increased magnitude of
wettability alteration was triggered. This prompted an increased desorption of the carboxylic
acid from the rock surface and SO2− 2−
4 occupied the freed site. The effectiveness of SO4 in
competing and displacing the carboxylic group of oil from the rock surface was explained by
Zhang et al. (2007). Tweheyo et al. (2006) observed similar trend during an imbibition and

123
An Experimental Investigation into the Impact of Sulfate. . .

Fig. 4 Contact angle with different exposing time at 95 ◦ C

contact angle experiments and reported that at high temperature, SO2−4 served as the crucial
active ion while Ca2+ and Mg2+ played a promotional role to the wettability alteration
process.
Although the reduction in contact angle was more prominent in SSB#4S compared to
SSB#8S, this was because the initial crude oil adsorption in SSB#4S test was not as strong as
that of SSB#8S. Another thing could be that SSB#4S offered the best and optimum wettability
alteration for all SO2−
4 concentration spiking because of an expected increased precipitation
of CaSO4 as SO2− 4 concentration increased. If this is the case, same trend was observed

123
A. Awolayo et al.

from the ζ -potential measurement, which showed that no surface charge alteration could be
achieved beyond SSB#4S. This is similar to what was identified by Tweheyo et al. (2006) in
chalk that a solubility limit of CaSO4 precipitation exists at a point and further increase in
SO2−
4 concentration becomes ineffective.

3.3 Coreflood and Ionic Analysis

Three waterfloods were conducted on the carbonate core plugs at similar experimental con-
ditions as discussed in Sect. 2.2 to study the ionic effect of smart brines on oil recovery. The
effluent brine was collected for ionic analysis of the active ions and non-active ions. The
results were normalized against the injected concentration to further confirm the postulated
mechanisms. Therefore, if no ionic exchange, rock dissolution or any other reaction occurred,
all normalized concentrations were expected to return to a value of 1.

3.3.1 Waterflood I

S#9 was saturated with reservoir oil to Swirr of 21.4 %. After the aging process, the core
was flooded sequentially with FB, SSB and SSB#0.5S. Altogether, 14.8 pore volumes (PV)
of FB were injected, and the recovery factor (RF) plateau of 75.6 % OOIC was reached
in about 11.5PV. The initial oil saturation was 78.6 % and decreased to 19.2 % after base
brine flooding. Injection halted when no more oil was produced from the core after 3PV,
the injection fluid was changed to SSB, and an incremental recovery factor (IRF) of 6.86 %
OOIC was documented. Thereafter, the injection fluid was switched to SSB#0.5S and no
additional oil recovery was observed after about 8.5PV. Injected smart water was observed
to reduce residual oil saturation (Sor ) by 5.39 %.
The displacement efficiency and differential pressure recorded are shown in Fig. 5. The
result indicated that at the same salinity/ionic strength while reducing the active ions concen-
trations, most especially SO2− 4 , no additional oil recovery was obtained. The pressure drop
profile showed that as the brine salinity decreased from FB to SSB#0.5, the differential pres-
sure relatively decreased and stabilized. This was because the injection fluid was switched
to brines of lower viscosity, which increased the mobility of the brine (Gupta et al. 2011).
The normalized effluent concentration profile is shown in Fig. 6 (active ions) and Fig. 7
(non-active ions). During FB injection, Ca2+ concentration remained high at 1.2 times and
Mg2+ concentration remained low at 0.9 times the FB slug, which revealed release of divalent
cations during aging process. While during cation replenishment at high temperature, Mg2+
displaced Ca2+ from the rock surface. The non-active ions settled below the injection level,
which indicated they replaced some ions at the diffuse layer. At this stage, the core was still
oil-wet, so the carboxylic group of the oil displaced SO2− 4 at the rock surface leading to its
increased amount in the effluent.
During SSB injection, concentration of Ca2+ , Na+ and Cl− increased and steeply declined
to almost the injection level, while the [Mg2+ ] was low and returned to the injection level in
about 5PV. The increased concentration of the non-active ions was due to the dilution between
FB and SSB. [SO2− 4 ] dropped low to half the injection slug concentration and increased back
to remain at about 1.2 times SSB slug. An ion exchange process occurred at the early stage
which involved the adsorption of SO2− 4 and co-adsorption of the divalent cations as proposed
by Zhang et al. (2007). For Ca2+ concentration not to return to the normalized concentration
of 1 and [Mg2+ ] at a low level indicated that an ion exchange process occurred, where Mg2+
displaced Ca2+ in the presence of SO2− 4 . This was responsible for the wettability alteration
process, hence the observed incremental recovery.

123
An Experimental Investigation into the Impact of Sulfate. . .

Recovery Differenal Pressure


100 1.7
90 IRF = 6.86% IRF = 0%
Displacement Efficiency % OOIC

80 RF = 75.57% 1.6

Differenal pressure MPa


70
1.5
60
50 1.4
40
1.3
30
20 FB SSB SSB#0.5S 1.2
10 Salinity: 261.2kppm Salinity: 43.3kppm Salinity: 43.3kppm
Ionic Strength: 5.19 Ionic Strength: 0.87 Ionic Strength: 0.87
0 1.1
0 5 10 15 20 25 30
Brine Injected PV
Fig. 5 Displacement efficiency and differential pressure versus brine injected PV—S#9

Ca2+ Mg2+ SO42-


10.0

FB SSB SSB#0.5S
C/Co

1.0

0.1
0 5 10 15 20 25 30
Brine Injected PV
Fig. 6 Normalized concentration of active ions in the effluent of Waterflood I

For SSB#0.5S, the non-active ion, Ca2+ and Mg2+ concentrations were constant in the
effluent due to similar concentration in both SSB and SSB#0.5S. As no change in ionic
interaction occurred except for SO2−
4 which remained high in the effluent and later returned
to the injection level due to the dilution of SSB with SSB#0.5S. This showed that SO2− 4 ,
which is considered vital in wettability alteration, was low in the injected brine. Therefore,
the rock wettability was not altered any further to generate additional recovery.
Waterflood II S#42 was made up of composite core plugs and was flooded sequentially
with FB, SSB and SSB#4S. The base brine FB recovered about 71.13 % OOIC. The injection
fluid was changed to SSB, which gave an incremental recovery of 5.7 % OOIC. The injection
fluid was replaced with SSB#4S and in about 3.5PV recorded an additional recovery of
3.09 % OOIC. The displacement efficiency and differential pressure plotted against injected
brine PV are presented in Fig. 8. The result showed that at the same salinity with increased
active ions concentration, most especially SO2−4 , higher incremental recovery and reduction
in Sor were obtained. The pressure drop profile also followed the expected trend as reported
in Waterflood I.

123
A. Awolayo et al.

Na+ Cl-
10.0

FB SB SSB#0.5S
C/Co

1.0

0.1
0 5 10 15 20 25 30
Brine Injected PV
Fig. 7 Normalized concentration of non-active ions in the effluent of Waterflood I

Recovery Differenal Pressure


100 3.5
90
Displacement Efficiency % OOIC

IRF = 5.7% IRF = 3.1% 3


RF = 71.1%

Differenal pressure MPa


80
70 2.5

60
2
50
1.5
40
30 1
20 FB SSB SSB#4S
Salinity: 261.2kppm Salinity: 43.3kppm 0.5
10 Salinity: 43.9kppm
Ionic Strength: 5.19 Ionic Strength: 0.87 Ionic Strength: 0.97
0 0
0 5 10 15 20 25
Brine Injected PV
Fig. 8 Displacement efficiency and differential pressure versus brine injected PV—S#42

The normalized effluent concentrations profile is shown in Fig. 9 (active ions) and Fig. 10
(non-active ions). During FB flooding, the concentrations of the active and non-active ions
in the effluent remained consistent with the injected brine level. During SSB injection, con-
centration of Ca2+ , Na+ and Cl− increased because of brine dilution and the non-active ions
were observed to return to the injected brine level. While Ca2+ concentration remained as
high as 3 times SSB slug, so were concentration of Mg2+ and SO2− 4 observed to drop below
the injected brine level and later returned. It was observed that the magnitude of the variation
in the concentration of SO2− + −
4 , Na and Cl observed here was lower compared to previous
case. This was as a result of differences in the rock characteristics, but similar multi-ion
exchange process was observed which elucidated the improved recovery observed as simi-
lar to the previous case. Just as the ions returned to the injected brine level (except Ca2+ ),
the displacement plateau was established as no other interactions occurred other than rock
dissolution (Hiorth et al. 2010), which occurred in the already swept zone.
The active ion concentrations in the effluent was observed to be lower than in the injected
slug during SSB#4S, while the non-active ions presence was more pronounced compared to
SSB flooding. This showed that SO2− 4 adsorbed more strongly with co-adsorption of Ca
2+ and

Mg2+ as the carboxylic components were detached. Similarly with high Cl− concentration

123
An Experimental Investigation into the Impact of Sulfate. . .

Ca2+ Mg2+ SO42-


10.0

C/Co

1.0

FB SSB SSB#4S
0.1
0 5 10 15 20 25
Brine Injected PV
Fig. 9 Normalized concentration of active ions in the effluent of Waterflood II

10.0 Na+ Cl-


C/Co

1.0

FB SSB SSB#4S
0.1
0 5 10 15 20 25
Brine Injected PV
Fig. 10 Normalized concentration of non-active ions in the effluent of Waterflood II

in the effluent, it showed that the active ions dominated the diffuse layer, which created better
access for the active ions to the rock surface. This changed the ion balance and the adsorbed
component altered the rock surface wettability. Similar behavior was seen during ζ -potential
and wettability-monitoring test where SSB#4S gave a better surface charge alteration and
more water-wetness compared to other brines.
Waterflood III S#313 was made up of composite core plugs and was flooded sequen-
tially with FB, SSB#2S, SSB#4S and SSB#8S. FB gave a recovery of 66.62 % OOIC,
and continuous injection of FB (1.5PV more) was to ensure the removal of all mobile oil,
which did not result in any production. The injection fluid was switched to SSB#2S, and
an incremental recovery of 2.89 % OOIC was documented. After the production seized, the
injection fluid was replaced with SSB#4S, which after 3.5PV recorded an additional recovery
of 5.71 % OOIC. Next, SSB#8S was injected and after 5PV, 1.11 % OOIC increment was
observed.
During FB flooding, Sor was observed to decrease to 26.59 % while smart brine injection
further decreased the Sor by 7.74 %. From Fig. 11, it can be inferred similarly to Waterflood
II that at the same salinity with increased SO2− 4 concentration, further incremental recovery
and reduction in Sor were obtained. This showed a good example of coreflood reproducibility.
The pressure drop profile as well showed a steady trend all through the flooding experiment
demonstrating that the injected fluids were relatively of similar viscosity.
The normalized effluent ion concentrations profile is presented in Fig. 12 (active ions)
and Fig. 13 (non-active ions). Similar to Waterflood I, during FB injection, the concentra-

123
A. Awolayo et al.

Recovery Differenal Pressure


100 0.800

90
IRF = 2.89% IRF = 1.11%
Displacement Efficiency % OOIC

IRF = 5.71%
80 RF = 66.6%

Differenal pressure MPa


0.750
70

60

50 0.700

40

30
0.650
20 FB SSB#2S SSB#4S SSB#8S
Salinity: 261.2kppm Salinity: 41.7kppm Salinity: 43.9kppm Salinity: 49.5kppm
10 Ionic Strength: 5.19 Ionic Strength: 0.88 Ionic Strength: 0.97 Ionic Strength: 1.14
0 0.600
0 5 10 15 20 25 30
Brine Injected PV
Fig. 11 Displacement efficiency and differential pressure versus brine injected PV—S#313

Ca2+ Mg2+ SO42-


10.0
C/Co

1.0

FB SSB#2S SSB#4S SSB#8S


0.1
0 5 10 15 20 25 30
Brine Injected PV
Fig. 12 Normalized concentration of active ions in the effluent of Waterflood III

tions of non-active ions in the effluent remained steady with the injected brine level, while
concentration of Ca2+ and Mg2+ remained low, which indicated that during aging process
there was release of divalent cations. After brine breakthrough, some Ca2+ and Mg2+ were
reabsorbed to the rock surface, which resulted in their low concentration in the effluent. Also
because of the oil-wet characteristics of the sample, the carboxylic component was observed
to displace SO2−
4 at the rock surface, and this led to its increased amount in the effluent.
Then during the SSB#2S injection, concentration of Ca2+ , Mg2+ , Na+ and Cl− increased
and steeply declined to the injected brine level except Mg2+ which went below and Ca2+
remained at twice the injected slug in about 5PV, whereas SO2− 4 concentration was observed
to drop settling below the injected concentration and later returned to the injected brine
level. Also during SSB#4S and SSB#8S flooding, SO2− 4 concentration in the effluent was
low and later returned to the injected brine level. Meanwhile, Mg2+ was relatively low
throughout and Ca2+ was found to level with the injected slug concentrations. The non-active
ion concentrations were high in the effluent as the SO2− 4 concentration dropped lower than

123
An Experimental Investigation into the Impact of Sulfate. . .

Na+ Cl-
10.0

C/Co

1.0

FB SSB#2S SSB#4S SSB#8S


0.1
0 5 10 15 20 25 30
Brine Injected PV
Fig. 13 Normalized concentration of non-active ions in the effluent of Waterflood III

the injection slug. Additional oil production recorded was due to improved water-wetness as
SO2−
4 adsorption to rock surface increased, with increased co-adsorption of Mg .
2+

The observation above complimented the ζ -potential and wettability-monitoring results


which showed that increasing SO2− 4 concentration led to surface charge alteration and hence
made the water film surrounding the rock more stable and thicker. This resulted in desorption
of oil from the rock surface and hence improved oil recovery. The incremental recovery
obtained in Waterflood III when SSB#8S was introduced after SSB#4S injection was ascribed
to no other reason than CaSO4 precipitation. This might have blocked some pore channels
and open up other channels that were not flooded previously to recover the residual oil.

3.4 Comparative Analysis

In the two composite flooding (Waterflood II and Waterflood III), the effectiveness of sulfate as
an active ion in facilitating wettability alteration toward more water-wetness was emphasized
compared to Waterflood I. The interactions between the active and non-active ions were more
pronounced when sulfate concentration was increased. This therefore led to reduction in the
residual oil saturation.
In the multi-ion exchange mechanism explained by Zhang et al. (2007), sulfate was
believed to compete with the carboxylic component of the oil to attach to the rock sur-
face and likewise the divalent cations to avoid charge imbalance. Therefore, this resulted
into desorption of the carboxylic components from the carbonate surface and hence changed
the rock wettability. The rate of ionic exchanges and the activity of Mg2+ were observed to
increase at higher temperatures (Kwak et al. 2014; Puntervold 2008). In this study, it was
observed that at the reservoir temperature of 110 ◦ C, more ionic exchanges occurred with
increased reactivity of Mg2+ .
Furthermore, multi-ion exchange still occurred at high brine salinity (lower compared to
formation water), which is similar with suggestion by RezaeiDoust et al. (2009). The authors
claimed that as long as the relative concentration of active and non-active ions differed
between the formation and invading smart brines, the multi-ion exchange could take place
at high invading brine salinity.
From the coreflood results discussed above, an improved recovery was observed due to
the alteration of brine ionic compositions, especially sulfate. In addition, the effluent ionic

123
A. Awolayo et al.

analysis results provided proof for the postulated mechanisms: multi-ion exchange and calcite
dissolution, whereas ζ -potential result showed that the alteration of surface charge was evi-
dent for improved water-wetness. Therefore, the rate of wettability alteration process depends
on the magnitude of surface charge alteration at the rock/brine and oil/brine interfaces. The
more similar the magnitude, the greater the repulsive forces. This resulted in wettability
alteration, which was as well supported by the results from contact angle measurement.

4 Conclusions

In this study, the ionic effect of smart water on enhancing oil recovery was thoroughly
investigated through series of experiment on candidate carbonate reservoirs. Based on the
experimental findings, while keeping in mind the limitations and assumptions made during
the study, the following conclusions are drawn:

• A direct relationship exists between increased sulfate concentration and ζ -potential of


the rock/brine interface. The degree of surface charge alteration is a function of the
magnitude of charge at the two interfaces. The more the sulfate concentration, the more
the two interfaces possess similar charges. This generated repulsive forces that caused
further alteration in surface charge.
• Experimental results demonstrate that increasing sulfate concentrations in smart brine
possess a dominant effect on displacement efficiency. The ultimate oil displacement
increased when sulfate concentration increased in smart brine relative to the base brine.
• Effluent concentration analysis presents evidence to support the postulated mechanism
responsible for the improved recovery.
• The wettability-monitoring study also demonstrated that increasing sulfate concentration
in smart brine is capable of altering the rock wetting state toward more water-wetness.
• Smart brine with four times sulfate concentrations proved to be the optimum sulfate
increment that could favorably improve oil recovery beyond any destructive impact.
• The observed mechanism responsible for the improved oil recovery is wettability alter-
ation (due to multi-ion exchange and surface charge alteration).

Acknowledgments The authors appreciate the financial support provided by The Petroleum Institute (PI)
in conducting this research work and the permission to present this paper. Authors are obliged to ADCO for
providing the experimental data and ZADCO UZFDS Laboratory for helping with effluent sample analysis.

References
Anderson, W.: Wettability literature survey-part 1: rock/oil/brine interactions and the effects of core handling
on wettability. J. Pet. Technol. 38(10), 1125–1144 (1986)
Austad, T.: Understanding of the EOR potential using smart water. In: Sheng, JJ. (ed.) Enhanced Oil Recovery
Field Case Studies, pp. 712. Gulf Professional Publishing: Waltham (2013)
Austad, T., Shariatpanahi, S., Strand, S., Black, C., Webb, K.: Conditions for a low-salinity enhanced oil
recovery (EOR) effect in carbonate oil reservoirs. Energy Fuels 26(1), 569–575 (2011)
Awolayo, A., AlSumaiti, A.M., Sarma, H.: An experimental study of smart waterflooding on fractured car-
bonate reservoirs. Paper presented at the ASME 2014 33rd international conference on ocean, offshore
and Arctic engineering, San Francisco, USA (2014a)
Awolayo, A.N., Sarma, H.K., Al-Sumaiti, A.M.: A laboratory study of ionic effect of smart water for enhancing
oil recovery in carbonate reservoirs. Paper presented at the SPE EOR conference at oil and gas west Asia,
Muscat, Oman (2014b)

123
An Experimental Investigation into the Impact of Sulfate. . .

Farooq, U., Tweheyo, M.T., Sjøblom, J., Øye, G.: Surface characterization of model, outcrop, and reservoir
samples in low salinity aqueous solutions. J. Dispers. Sci. Technol. 32(4), 519–531 (2011)
Gupta, R., Smith Jr. P., Willingham, T., Lo Cascia, M., Shyeh, J., Harris, C.: Enhanced waterflood for middle
east carbonate cores—impact of injection water composition. Paper presented at the SPE middle east oil
and gas show and conference, Manama, Bahrain (2011)
Hiorth, A., Cathles, L., Madland, M.: The impact of pore water chemistry on carbonate surface charge and oil
wettability. Transp. Porous Media 85(1), 1–21 (2010)
Høgnesen, E., Strand, S., Austad, T.: Waterflooding of preferential oil-wet carbonates: Oil recovery related
to reservoir temperature and brine composition. Paper presented at the SPE Europec/EAGE annual
conference, Madrid, Spain (2005)
Jadhunandan, P., Morrow, N.: Effect of wettability on waterflood recovery for crude-oil/brine/rock systems.
SPE Reserv. Eng. 10(1), 40–46 (1995)
Kwak, H.T., Yousef, A.A., Al-Saleh, S.: New insights on the role of multivalent ions in water-carbonate rock
interactions. Paper presented at the SPE improved oil recovery symposium, Tulsa, Oklahoma, USA
(2014)
Lager, A., Webb, K.J., Black, C., Singleton, M., Sorbie, K.: Low salinity oil recovery—an experimental
investigation1. Petrophysics 49(1), 28–35 (2008)
Lager, A., Webb, K.J., Black, C.J.J.: Impact of brine chemistry on oil recovery. Paper presented at the 14th
European symposium on improved oil recovery, Cairo, Egypt (2007)
Ligthelm, D., Gronsveld, J., Hofman, J., Brussee, N., Marcelis, F., van der Linde, H.: Novel waterflood-
ing strategy by manipulation of injection brine composition. Paper presented at the EUROPEC/EAGE
conference and exhibition, Amsterdam, The Netherlands (2009)
McGuire, P., Chatham, J., Paskvan, F., Sommer, D., Carini, F.: Low salinity oil recovery: an exciting new
EOR opportunity for Alaska’s north slope. Paper presented at the SPE western regional meeting, Irvine,
California (2005)
Pierre, A., Lamarche, J., Mercier, R., Foissy, A., Persello, J.: Calcium as potential determining ion in aqueous
calcite suspensions. J. Dispers. Sci. Technol. 11(6), 611–635 (1990)
Puntervold, T.: Petroleum Engineering. University of Stavanger, Norway (2008)
RezaeiDoust, A., Puntervold, T., Strand, S., Austad, T.: Smart water as wettability modifier in carbonate
and sandstone: a discussion of similarities/differences in the chemical mechanisms. Energy Fuels 23(9),
4479–4485 (2009)
Rod, S.: Quantification of uncertainty in recovery efficiency predictions: lessons learned from 250 mature
carbonate fields. Paper presented at the SPE annual technical conference and exhibition, Denver, Colorado
(2003)
Sheng, J.: Enhanced Oil Recovery Field Case Studies. Gulf Professional Publishing, Houston (2013a)
Sheng, J.J.: Review of surfactant enhanced oil recovery in carbonate reservoirs. Adv. Pet. Explor. Dev. 6(1),
1–10 (2013b)
Strand, S., Austad, T., Puntervold, T., Høgnesen, E.J., Olsen, M., Barstad, S.M.F.: Smart water for oil recovery
from fractured limestone: a preliminary study. Energy Fuels 22(5), 3126–3133 (2008)
Strand, S., Høgnesen, E.J., Austad, T.: Wettability alteration of carbonates—effects of potential determining
ions (Ca2+ and SO2− 4 ) and temperature. Colloids Surf. A Physicochem. Eng. Asp. 275(1), 1–10 (2006)
Takamura, K., Chow, R.S.: The electric properties of the bitumen/water interface Part II. Application of the
ionizable surface-group model. Colloids Surf. 15, 35–48 (1985)
Tang, G.Q., Morrow, N.R.: Influence of brine composition and fines migration on crude oil/brine/rock inter-
actions and oil recovery. J. Pet. Sci. Eng. 24(2), 99–111 (1999)
Tweheyo, M., Zhang, P., Austad, T.: The effects of temperature and potential determining ions present in
seawater on oil recovery from fractured carbonates. Paper presented at the SPE/DOE symposium on
improved oil recovery, Tulsa, Oklahoma, USA (2006)
Webb, K., Black, C., Al-Ajeel, H.: Low salinity oil recovery-log-inject-log. Paper presented at the SPE/DOE
symposium on improved oil recovery, Tulsa, Oklahoma, USA (2004)
Yi, Z., Sarma, H.: Improving waterflood recovery efficiency in carbonate reservoirs through salinity variations
and ionic exchanges: a promising low-cost smart-waterflood approach. Paper presented at the Abu Dhabi
international petroleum conference and exhibition, Abu Dhabi, UAE (2012)
Yousef, A., Al-Saleh, S., Al-Kaabi, A., Al-Jawfi, M.: Laboratory investigation of novel oil recovery method
for carbonate reservoirs. Paper presented at the Canadian unconventional resources and international
petroleum conference, Calgary, Alberta, Canada (2010)
Zahid, A., Shapiro, A., Skauge, A.: Experimental studies of low salinity water flooding in carbonate reservoirs:
a new promising approach. Paper presented at the SPE EOR conference at oil and gas West Asia, Muscat,
Oman (2012)

123
A. Awolayo et al.

Zhang, P., Austad, T.: Wettability and oil recovery from carbonates: effects of temperature and potential
determining ions. Colloids Surf. A Physicochem. Eng. Asp. 279(1), 179–187 (2006)
Zhang, P., Tweheyo, M.T., Austad, T.: Wettability alteration and improved oil recovery in chalk: the effect of
calcium in the presence of sulfate. Energy Fuels 20(5), 2056–2062 (2006)
Zhang, P., Tweheyo, M.T., Austad, T.: Wettability alteration and improved oil recovery by spontaneous imbi-
bition of seawater into chalk: impact of the potential determining ions Ca2+ , Mg2+ , and SO2−
4 . Colloids
Surf. A Physicochem. Eng. Asp. 301(1), 199–208 (2007)

123

You might also like