You are on page 1of 8

Journal of Molecular Liquids 250 (2018) 396–403

Contents lists available at ScienceDirect

Journal of Molecular Liquids

journal homepage: www.elsevier.com/locate/molliq

Drivers of low salinity effect in sandstone reservoirs


Ehsan Pooryousefy, Quan Xie ⁎, Yongqiang Chen, Ahmad Sari, Ali Saeedi
Department of Petroleum Engineering, Curtin University, 26 Dick Perry Avenue, 6151 Kensington, Western Australia, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Wettability of oil/brine/rock system is a fundamental petro-physical parameter, governing the subsurface multi-
Received 22 October 2017 phase flow behaviour, thus affecting hydrocarbon recovery. However, understanding the controlling factor(s)
Received in revised form 25 November 2017 which drives the process is incomplete. Therefore, a detailed examination and characterization of reservoir sur-
Accepted 28 November 2017
face forces as well as pore-surface chemistry are of vital importance. We combined contact angle, zeta potential
Available online 8 December 2017
and surface complexation modelling to demonstrate that DLVO theory and surface complexation modelling pre-
Keywords:
dict the same wettability trends. Moreover, we introduced Z* parameter (Z∗ = |Zoil/brine + Zbrine/rock |) to manage
Low salinity water flooding low salinity water flooding, thus mitigate the uncertainty of low salinity effect.
Wettability © 2017 Elsevier B.V. All rights reserved.
Contact angle
Zeta potential
Surface complexation modelling

1. Introduction chemistry to measured zeta potential [26,27]. However, the factors con-
trolling the wettability in oil/brine/rock system is yet to be clearly ex-
Wettability of oil/brine/rock system governs subsurface multiphase plained, thus limiting the implementation of low salinity water
flow behaviour and saturation distribution of the fluids, thus affecting flooding in oil fields. For instance, while most reports suggest that the
oil recovery from hydrocarbon reservoirs [1,2]. Water-flooding tech- presence of divalent cations, clays, and acidic compounds from crude
nique has been widely used to maintain reservoir pressure, thereby im- oils, are of vital importance to yield wettability transition towards
proving oil recovery [3]. Recent studies [4–6] proposed a novel means to more water-wet [8,16], Hassenkam et al. [28] found the opposite, argu-
enhance oil recovery by manipulating the composition of injected water ing that a reduction in adhesion force of the oil/brine/rock system can be
chemistries, in a process called LoSal flooding by BP [7,8], Smart Water achieved even from a pure quartz surface with a model oil composed
flooding by its originators, Austad and co-workers at the University of only of CH3. To date, uncertainties about the nature of the physics con-
Stavanger, Saudi Aramco [9], and Designer Water flooding by Shell [10, trolling the interaction of the oil/brine/rock system remain, with some
11]. reports suggest using salinity level to interpolate between the relative
Several mechanisms was proposed to decipher what factors control permeability curves [29,30], whereas others suggest using individual
the low salinity effect: fines mobilization [12], limited release of mixed- ion, e.g., Ca2+, Mg2 + and SO2− 4 in the solution [31,32]. The nature of
wet particles [12], increased pH and reduced IFT similar to the alkaline the wettability in oil/brine/rock system remains something of a mys-
flooding [13], multi-component ion exchange (MIE) [8,14–16], expan- tery, with many questions about its chemical nature at a molecular
sion of the electrical double layer [17], salt-in effect [18], salting-out ef- level.
fect [19] and osmotic pressure [20]. To better model and manage the low salinity water flooding, we
However, the main mechanism behind low salinity water flooding is hypothesised that zeta potential at the interfaces of oil/brine and
thought to be wettability alteration [21,22]. Yet, knowing what specific brine/rock govern the wettability transition because zeta potential is a
factors govern the reservoir wettability in the presence of crude oil and reflection of geo-chemical reactions and physisorption in oil/brine/
formation brine has long been a goal of reservoir engineers. To under- rock system. To test this hypothesis, we measured the contact angle of
stand the factors controlling the interplay between fluid-fluid and oil droplets on the surface of muscovite sheets in the presence of CaCl2
fluid-rock interfaces, atomic force microscopy (AFM) [23,24] and quartz with various salinity levels (2000, 10,000, 30,000 and 50,000 mg/l), at
crystal microbalance with dissipation (QCM-D) [25] were used to inves- temperatures (50 and 100 °C). We also measured the zeta potential of
tigate the effect of water chemistry on adhesion force at nano-scale. In oil/brines and brines/muscovite. Moreover, we did a geochemical
addition, surface complexation modelling (SCM) was also applied to un- study to decipher the factors governing the wetting characteristics of
derstand the wettability of reservoir rocks to relate the surface the oil/brine/muscovite system. Furthermore, to correctly predict the
performance of the low salinity water flooding, we proposed a new pa-
⁎ Corresponding author. rameter (Z*), which is the absolute sum of the zeta potential of oil/brine,
E-mail address: quan.xie@curtin.edu.au (Q. Xie). and brine/rock, to correlate with contact angle, thus system wettability.

https://doi.org/10.1016/j.molliq.2017.11.170
0167-7322/© 2017 Elsevier B.V. All rights reserved.
E. Pooryousefy et al. / Journal of Molecular Liquids 250 (2018) 396–403 397

2. Methods 2.1.3. Crude oil


Texas crude oil with acid number 1.58 mg KOH/g and base number
2.1. Materials 1.02 mg KOH/g was used in this work to measure the contact angle on
the muscovite substrate in the presence of the various salinity level.
2.1.1. Substrates The density of the experimental oil was 0.89 mg/ml.
To examine the effect of water chemistry on the contact angle of oil
droplet on the surface of minerals, we used muscovite sheet (supplied 2.2. Experimental procedure
by Wards Scientific) as substrates. This is due to muscovite being the
main constituent of different formed minerals in sedimentary rocks Contact angle measurements were conducted using high pressure/
[33]. In addition, muscovite represents the common clay minerals in high temperature IFT 700 cell (Fig. 2). This apparatus is equipped with
sandstone reservoirs and has a similar chemical structure to Illite [34]. a high resolution camera and a built-in contact angle measurement soft-
Furthermore, muscovite has a cation exchange capacity 3–4 times ware enabling the constant monitoring of contact angle alteration. The
greater than kaolinite [35], which is believed to be important to exhibit treated substrates were placed into the IFT 700 cell and vacuumed for
low salinity EOR effect at reservoirs [6,36]. 1 h until the pressure stabilised at 0.1 bar. The degassed solution was de-
To avoid the effect of surface roughness on the contact angle, we ployed into the cell with pressure and temperature adjusted to 200 bar
used the atomic force microscopy (AFM) (WITec, ALPHA 300 RA for at 50 °C. Subsequently, the oil droplet was released on the substrate
combined Raman-AFM imaging) to examine the surface roughness of from a capillary needle with diameter of 0.06 mm. In order to obtain a
the muscovite substrate after cleaving prior to contact angle measure- reliable contact angle measurement, the contact angle was monitored
ments. AFM test shows that the surface roughness of the muscovite for a period of 72 h. The standard deviation for contact angle measure-
was in a range of 0–2.5 nm (Fig. 1), implying that the effect of roughness ments was ±2°.
on contact angle test is negligible because the surface roughness is
much less than the thickness of the water film at the surface [14]. 2.2.1. Zeta potential measurements
Note: to obtain a clean surface, the muscovite substrates was simply Surface charges would be generated at the interfaces of fluid-fluid
cleaved after each experiment. and fluid-rock due to the ionization of the chemical group in the pres-
ence of aqueous ionic solutions [43]. The strength of electrostatic
charges at interfaces can be determined by zeta potential, which may
2.1.2. Ionic aqueous solutions control the oil/brine/rock system wettability. We thus measured the
To test our hypothesis, we used CaCl2 with various concentrations zeta potential of interface of oil/brine and brine/muscovite at various sa-
to examine the contact angle of oil/brine/rock system. This is because linity level using Zetasizer Nano ZS manufactured by Malvern.
Ca2 + is a major divalent cation component in formation brines, and Due to the fact that measured zeta potentials are much more stable
Ca 2 + ions can act as bridges to bond the polarized ends from the at 25 °C than high temperatures, we performed all zeta potential mea-
crude oil to the surface of pore wall [35,37]. Furthermore, Ca2 + has surements at 25 °C to examine the surface potential dependencies of
been identified as one of the main PDI (potential determining ions) oil and rock surface reactivity. Note: at higher temperatures, zeta poten-
which has been previously confirmed to be the main contributors tials tend to decrease uniformly with an increase in temperature [44].
to low salinity effect compared to monovalent cations such as Na+ The experimental procedures to test zeta potential follow the proce-
[38,39]. Meanwhile, previous studies has also confirmed that Mg2 + dures proposed by Xie et al. [45]. The standard deviation for zeta poten-
has a similar behaviour to Ca2 + [40,41]. However, the affinity of tial measurements was ±1 mV.
Ca2 + towards clay surfaces is stronger than that of Mg2 +. We there-
fore used CaCl2 at different salinity level (2000, 10,000, 30,000 and 3. Results and discussion
50,000 mg/l) to test zeta potential and contact angle in oil/brine/
muscovite system. 3.1. Effect of salinity and temperature on contact angle
Given that pH strongly affects oil and mineral surface chemistry,
which consequently changes the zeta potential and the physisorption Water contact angle of oil droplets on muscovite surface decreased
[42,43], we manually adjusted the pH of the ionic aqueous solutions to with increasing salinity, implying that the system wettability was
7 using HCl and NaOH standard solutions before aqueous solutions shifted from slightly oil-wet to slightly water-wet with increasing salin-
were degassed in the desiccator for a period of 24 h. ity. (Note: all final contact angle results were reported after the

Fig. 1. AFM topographic image of the cleaved muscovite substrate used for contact angle measurements.
398 E. Pooryousefy et al. / Journal of Molecular Liquids 250 (2018) 396–403

Valve Hand pump


(brine)
High definition camera

Light source

HT-HP IFT cell


Valve Hand pump
(crude oil)

Fig. 2. Schematic diagram of the high pressure-high temperature IFT cell for contact angle measurements.

equilibrium was reached, and the standard deviation for contact angle muscovite due to the opposite polarity of the zeta potential of oil/
measurements was ±4°.). For example, at temperature 100 °C, the con- brine and brine/muscovite. This triggers oil-wet characteristics because
tact angle was 118° in the presence of 2000 mg/l CaCl2. However, the the Van-der-Waal force is always attractive. Yet, the 50,000 mg/l CaCl2
contact angle decreased from 118° to 102° when salinity increased to yielded a positive zeta potential for both oil/brine (14.73 mV) and
10,000 mg/l. Further increasing the ionic concentration to 30,000 mg/l brine/muscovite (6.92 mV), thus triggering positive double layer expan-
decreased the contact angle to 89°, and the smallest contact angle was sion forces [47,48]. The positive double expansion forces in return com-
recorded when the salinity increased to 50,000 mg/l. Together, the re- pensate the Van-der-Waal attractive force, thus likely shifting the
sults show that low concentration of Ca2+ triggered an oil-wet system. disjoining pressure from negative to positive [49]. Note: a negative
Yet, the high concentration of Ca2+ led to a slightly water-wet system. disjoining pressure means attractive forces, indicating an oil-wet sys-
The similar results were reported by Mohsenzadeh et al. [46] who tem. Rather, a positive disjoining pressure means repulsive forces, indi-
found that successive dilution of formation brine (FB) increased the cating a water-wet system [50].
contact angle from 63° in formation brine (203,000 ppm), to 130° in While multicomponent ion exchange (MIE) [51] has long been con-
20 times diluted formation brine, altering wettability from water-wet sidered as one of the main mechanisms of low salinity water flooding,
to oil-wet. This may explain why the low salinity water flooding is not our wettability alteration results on Texas crude, did not support this
always observed using the conventional dilution approach. theory. Lager et al. [16] discussed four main mechanisms of adsorption
Contact angle also decreased slightly with decreasing temperature, of organic matter to the surface of rock (former proposed by Sposito
but temperature played a minor role in the wetting characteristics of [52]) and hypothesised that removing divalent cations, such as Ca2+
the muscovite surface compared with salinity of Ca2+ (Fig. 3). For in- and Mg2+ from the solution, would break the bonds between the oil
stance, the contact angle decreased by 2-5° as the temperature dropped and the rock surface, which eventually dissociate the crude oil from
to 50 °C from 100 °C. However, the salinity variation rendered 35° de- the rock surface, leading to a more water-wet surface. They also
crease with increasing Ca2+. The similar trend were also reported by
Xie et al. [42] Who found that the contact angles increased with temper-
ature for all brines (formation brine, softened brine and 100 times dilut-
ed formation brine), but the effect of temperature on the contact angle
was not as pronounced as that of water chemistry. They observed that
the contact angle varied from 39° to 42° with increasing temperature
from 60 to 140 °C in the presence of softened brine.

3.2. Effect of salinity on zeta potential

To gain a deeper understanding of the results from the contact angel


tests, we examined the effect of salinity on zeta potential of fluid-fluid
and fluid-rock. Our results show that zeta potential of oil/brine in-
creased from − 2.82 to 14.73 mV as salinity increasing from 2000 to
50,000 mg/l, suggesting a significant adsorption of the Ca2+ at the oil/
brine interface due to the presence of huge amount of Ca2+ in high sa-
linity level. Zeta potential of brine/muscovite increased from 4.96 to
6.92 mV with increasing salinity, showing that effect of Ca2 + on the
zeta potential of brine/muscovite was not as pronounced as the inter-
face of oil/brine. Together, 2000 mg/l CaCl2 generated an attractive dou- Fig. 3. Contact angle vs. salinity level (CaCl2) at temperature of 50 and 100 °C at constant
ble layer force (collapse of thin water film) in the system of oil/brine/ pH 7.
E. Pooryousefy et al. / Journal of Molecular Liquids 250 (2018) 396–403 399

observed a higher oil recovery by removing the Ca2+ and Mg2+ from Table 1
the solution compared to the injection of a high concentration of NaCl Surface complexation model input parameters [27].

solution [16]. In contrast to MIE theory, we observed a shift from oil- Reaction Log K25oC
wet to slightly water-wet with increasing concentrations of Ca2+. This −NH+ = −N + H+ −6.0
discrepancy is very likely attributed to the surface chemistry of crude −COOH = –COO− + H+ −5.0
oil and minerals. For example, low concentration of the polarized ends −COOH + Ca2+ = −COOCa+ + H+ −3.8
(COO−, –NH+, and –COOCa+) from crude oil (e.g., low acid number NNa + Ca2+ ≥ Ca + 2Na+ 0.21
NNa + H+ ≥ H + Na+ 4.6
and base number) may not trigger the shift of the zeta potential of oil/
brine from negative to positive with increasing salinity, particularly in-
crease in concentrations of divalent cations, e.g., Ca2+, Mg2+. This is due
to a lack of –COOCa+ at the interface of oil/brine. Moreover, the mineral the oil/brine/rock interfaces (Note: the surface species were calculated
surface chemistry is an essential factor to cause this discrepancy. For ex- by PHREEQC version 3.30 (Parkhurst and Appelo 2013)).
ample, as pointed out by Brady et al. [53], MIE may work for the min- Where the “N” denotes negatively charged site on muscovite surface
erals with surface charges governed by edge charges, like kaolinite, while the “−” denotes negative charged site on oil surface. When the oil
rather than minerals controlled by basal plane charges, e.g., muscovite, adheres to the rock surface, ion exchange occurs, which can be
illite, chlorite, smectite. This also explains why Seccombe et al. [15] ob- expressed as the chemical reaction: NNa + − NH− ≥ − NH + Na+,
served a shift of wettability from slightly water-wet to strongly water- and NNa + −COOCa+ ≥ −COOCa + Na+.
wet in Endicott Field with high content of kaolinite with decreasing sa-
linity. They also reported that improvement in incremental oil recovery 3.4. Effect of salinity on number of oil/brine surface species
with decreasing salinity is proportional to the kaolinite content in the
reservoir. Similar to Sccombe et al.'s report [15], Xie et al. [42] also ob- Salinity strongly affected the number of oil/brine surface species,
served a wettability (rock is rich in kaolinite) shift from slightly named site density, particularly at the pH values lower than 7.5 (Fig.
water-wet to strongly water-wet in high temperature (140 °C) with de- 5). For example, the concentration of the –COO− and –NH+ at the inter-
creasing salinity. Therefore, we believe that decreasing salinity and re- face decreased with increasing salinity at a given pH level, although the
moving divalent cations (Ca2+ and Mg2+) would trigger low salinity magnitude of the decrease of –NH+ was greater than –COO−. Yet, the
EOR-effect in kaolinite-bearing sandstone reservoirs. This is because site density of –COOCa+ increased with increasing salinity. Together,
the surface charge of kaolinite-bearing sandstone is governed by edge these imply that the oil surface charge is likely to be dominated by –
charges which controls the surface complexation of oil/brine/rock sys- COOCa+ with the increasing Ca2+, particularly for crude oil with high
tem [43]. However, in this context, for muscovite or a mineral with acid number. This is because that more carboxyl groups (–COO−) will
the same structure-bearing as sandstone, decreasing salinity level diffuse from the interior to the surface at pH N 5, but these –COO−
would yield oil-wet wetting characteristic of the system of oil/brine/ would not be stable in the presence of Ca2 + and will be easily trans-
rock, thus hindering low salinity EOR-effect. This may explain the ques- formed to –COOCa+ [43].
tion: why did the conventional dilution approach do not trigger incre- pH played a great role in the site density of oil/brine at a given salin-
mental oil recovery in high temperature for the reservoir with a ity. For example, the site density of both –NH+ and –COO− decreased
certain amount of illite and crude oil with low acid number (0.25 mg with increasing pH, particularly at pH b 7.5. Moreover, the site density
KHO/g) [54]. Similar experimental results were also observed by of –NH+ declined much faster than that of –COO– because the depro-
Brady et al. [55] who reported that the oil adsorption in the middle tonation of –NH+ is more active than that of –COO− at low pH level.
Bakken with clay mainly composed of illite, decreased with increasing Yet, the surface site density of all of the surface chemical groups
salinity level (NaCl) at a given pH, suggesting that decreasing salinity remained constant at pH N 8. Also, at this range of pH level, the effect
shifts the wettability of oil/brine/rock system from water-wet to oil- of salinity on the surface site density was negligible (Fig. 5).
wet surface, in line with this work. While the wettability of oil/brine/ Surface complexation modelling show that the site density of the
rock system strongly relates to the surface chemistry of oil/brine and surface chemical groups strongly correlated with the zeta potential of
brine/rock, zeta potential of the fluid-fluid, and fluid-rock is believed oil/brine in the presence of the various salinity level. For instance, at
to be the parameter that correlates surface chemistry to physical ad- pH = 7, the site density of the –COOCa+ increased from 14.76 to
sorption and gives a reasonable prediction of low salinity water 19.55 mmol/m2 as the salinity level increased from 2000 to
flooding. Note: in this work, pH was kept constant which might be the 50,000 mg/l. Yet, the site density of the –COO− and –NH+ decreased
reason of why we observed the trend: increasing salinity triggers a from 20.90 and 4.98 to 18.39 and 0.66 mmol/m2 as the salinity level in-
more water-wet system. creased from 2000 to 50,000 mg/l, respectively. Together, the increase
of the –COOCa+ at the interface of oil/brine caused a shift in the zeta po-
tential from negative to positive as the salinity increased from 2000 to
3.3. Surface complexation modelling 50,000 mg/l. This was supported by the zeta potential results, showing
that the zeta potential of oil/brine increased from −2.82 to 14.73 mV
In this context, a surface complexation model developed by Brady et as the salinity increased from 2000 to 50,000 mg/l (Fig. 4). Note: the in-
al. [27,55] was applied to identify the individual charged species that crease of the Na+ in the aqueous solution would decrease the site den-
give rise to oil and rock zeta potentials, and then used to calculate the sity of –COOCa+ at a given pH [43] largely because of decreasing the
degree of electrostatic bridges between the oil and rock. DLVO theory activity coefficient of dissolved Ca2 +. Brady et al. [43] reported that
and surface complexation modelling both recognize an electric double the surface concentration of –COOCa+ decreased with increasing NaCl.
layer and the existence of charged surface species whose concentrations Yet, the surface concentration of –COO− increased with increasing the
depend upon the chemical composition of the water, oil and minerals concentration of the NaCl. This imply that negative zeta potential likely
[42]. In the surface complexation model, the muscovite surface area be observed in the presence of high salinity level but with low concen-
was assumed to be 100 m2/g with site density of 10 μmol/m2. The chem- tration of the Ca2+.
ical reactions on the surface of muscovite and crude oil can be described The variation of the site density of the surface chemical groups in the
by the following reactions (Table 1). For comparison with the zeta po- presence of aqueous ionic solutions likely alters the interfacial tension of
tential results obtained at 25 °C, we used the Log K at 25 °C (Table 1) the oil/brine. For example, Tang and Morrow [12] reported that the IFT
for all of the chemical reactions as temperature has a minor effect [42] decreased from 20 to 16 mN/m from high to low salinity brine. This is
and water chemistry dominated the surface complexation species at attributed to the increase of pH, which governs the oil/water surface
400 E. Pooryousefy et al. / Journal of Molecular Liquids 250 (2018) 396–403

Given that the surface chemistry of the experimental oil consists of –


COO-, −COOCa+ and –NH+, the electrostatic bridge of −COOCa+ and –
NH+ were summed up to show the total electrostatic bridges at the sur-
face of muscovite in the presence of different salinity level of CaCl2. Re-
sults show that at pH = 7, the electrostatic bridges decreased with
increasing salinity, implying the wettability of the system is shifting to-
wards water-wet, consistent with the contact angle test (Fig. 3). Fig. 7
also shows that the total electrostatic bridges was strongly affected by
pH apart from the salinity. For example, at pH b 6, the total electrostatic
bridges increased with increasing salinity, suggesting that with the
presence of mineral surface like muscovite, the wettability of rock sur-
face may shift towards more water-wet with decreasing salinity. How-
ever, at pH N 7, decreasing salinity may generate oil-wet surface. This
may explain that why the low salinity EOR-effect is not always observed
using conventional dilution approach.

Fig. 4. Zeta potential of oil/brine and brine/muscovite in the presence of different 3.6. Implications
concentration of CaCl2.

Understanding the governing factor(s) and mechanism(s) of the low


charge [56,57]. However, due to the limited reduction of interfacial ten- salinity water flooding in sandstone reservoirs is of fundamental impor-
sion, the decrease in capillary number with low salinity water is negligi- tance to constrain the intrinsic uncertainties in implementation of low
ble to cause chemical water EOR effect [12,58]. salinity water flooding technique. This study provides insights into the
interaction of interfaces of fluid-fluid and fluid-rock, and provided evi-
dence relating the zeta potential of interface of oil/brine and brine/
3.5. Effect of salinity on the electrostatic bridges rock to wettability. We also demonstrate that the surface complexation
modelling essentially contributes to a deeper understating of the factors
Similar to the oil surface, the salinity level strongly affected the elec- controlling low salinity effect in sandstone reservoirs at the interfacial
trostatic bridges for oil-muscovite. For example, the electrostatic brid- scale (nano-scale).
ges of N−NH decreased with increasing salinity at a given pH. Fig. 6 Moreover, given that charged surfaces, e.g. negative surface in posi-
shows that at pH = 7, the electrostatic bridges of N−NH decreased tively charged solution, the brine chemistry, e.g., salinity and ions
from 2.68 to 0.66 μmol/m2 as the salinity increased from 2000 to strength, dominate the Debye-length and zeta potential [59], decreasing
50,000 mg/l. This is because the exchangeable Na+ embedded into the the salinity and ionic strength causes the zeta potential of fluid-fluid and
muscovite was exchanged by Ca2+ with increasing the concentration fluid-rock to become strongly negative [60], thus resulting in a larger
of Ca2+. In return, the chemical equilibrium of NNa + −NH− ≥ − NH double layer expansion force, which compensates the Van-der-Waal at-
+ Na+ shifted to left, thus decreasing N− NH. This implies that the tractive force, thereby shifting the reservoir wettability towards more
rock surface likely becomes more water-wet as increasing salinity for water-wet [49]. We therefore particularly focused on the electrical dou-
the system with basal charged minerals (e.g., illite, smectite and chlo- ble layer force, assuming that the geometry of oil and rock surfaces is
rite), particularly for the crude oil with high base number. simplified as flat-charged surfaces with different potentials separated
Fig. 6 also shows that the number of electrostatic bridges (N− by a thin film of electrolyte [61]. Considering that the same polarity of
COOCa) increased with increasing salinity. For example, N−COOCa in- the zeta potential of oil/brine and brine/rock triggers repulsive electrical
creased to 9.27 from 7.26 μmol/m2 with increasing salinity from 2000 double layer force, and the opposite polarity of the zeta potential of oil/
to 50,000 mg/, because the site density of –COOCa+ increased with in- brine and brine/rock generates the attractive electrical double layer
creasing salinity (Fig. 5), enabling the chemical equilibrium (NNa + force [1], we proposed a new parameter, Z*, which is the absolute
−COOCa+ ≥ −COOCa + Na+) to shift towards right-hand side. This im- value of zeta potential of oil/brine and brine/rock (Eq. 1). Note: Z* can
plies that the oil/brine/rock system likely becomes more oil-wet with be applied with both opposite, and same polarity of zeta potential at
increasing salinity. This might explain why the low salinity EOR effect the interface of fluid-fluid and fluid-rock. A plot of Z* versus contact
was observed using oil with high acid number [42]. angle for various references [42,45] was then constructed (Table 2 and

Fig. 5. Site density of the surface chemical group vs pH at the interface of the oil/brine in the presence of various salinity level.
E. Pooryousefy et al. / Journal of Molecular Liquids 250 (2018) 396–403 401

Fig. 6. Electrostatic bridges vs pH of surface species in the system of oil/brine/muscovite.

Fig. 8). sandstone used by Nasralla et al. were rich in clays (illite = 2.7 wt%,
and kaolinite = 3.7 wt%), we assumed that the contact angle of oil on
Z  ¼ Zetaoil=brine þ Zetabrine=rock ð1Þ the mica substrate approximately represent the contact angle of oil on
the sandstone substrates in the presence of the same brines.
We demonstrate that the wettability of the oil/brine/rock system Together, although zeta potential was proposed to interpret the low
can be correlated with the zeta potential of fluid-fluid and fluid-rock. salinity effect [17,45,60,62–64], we believe this is the first work to ex-
For example, we show that the contact angle decreased linearly with in- plicitly relate zeta potential of oil/brine and brine/rock to contact
creasing the Z* for various oil/brine/rock system [42,45], implying that angle using Z*, thus providing a framework to design injected water
wettability of the system shifts towards more water-wet due to the chemistry. Note: oil composition, to be more specific, the acid number
same polarity of the zeta potential of fluid-fluid and fluid-rock. More- and base number, affects number and type (− NH+, − COO−, and –
over, we found that the slope of the function of zeta potential versus COOCa+) of the surface species, thereby the zeta potential of oil/brine
Z* is almost the same, implying that the wettability of oil/brine/system [27,65].
follows the same trend with change of Z* regardless of the water chem- In general, to constrain the intrinsic uncertainty of low salinity water
istries, composition of crude oil and mineral surface chemistry. This flooding in field, we therefore proposed Z* (Eq. 1), relating the zeta po-
demonstrates that Z* should be used as a strong indicator to reflect tential of the oil/brine/rock system to the wetting characteristic of the
the wettability shift of the oil/brine/rock system. Meanwhile, we also reservoir rock. We argue that the new parameter, Z*, may be used to in-
observed that the intercept at the Y axis varies with various system of terpolate between the bounding curves (relative permeability curves
oil/brine/rock. We believe that this is attributed to the nature of the from formation water and low salinity water), correctly simulating the
given oil/brine/rock system, which is again also related to the surface low salinity water flooding performance rather than using the salinity
complexation and physisorption although a more quantitative analysis level and individual ion (e.g., Ca2+, SO2− 4 etc.). This is because the Z*
remains to be made. of the oil/brine/rock system is the physical reflection of the salinity
To the best of our knowledge, while zeta potential of oil/brine and level, ion type, pH, oil and mineral surface chemistry [17,26,60]. Note
brine/minerals have been widely measured to understand the wettabil- that to gain a complete understanding of oil/brine/rock system, we be-
ity alteration induced by low salinity water, we did not find any litera- lieve that both electrostatic (double layer expansion) and non-electro-
ture reporting the zeta potential of oil/brine and brine/rock, and static physisorption (van der waals forces) together with competitive
contact angle using the same brine and rock. Nasralla and Nasr-El-Din ion chemisorption (ion exchange and surface complexation) govern
[17] measured zeta potential of oil/brines, brines/sandstone, but the the interaction of oil/brine/rock system, thus wettability. To understand
contact angle was measured at the surface of mica. Given that the the relative contribution of double layer expansion and surface com-
plexation, we believe that the mechanisms outlined above need to be
incorporated into total surface energy [66] which needs to be investigat-
ed in future.

Fig. 7. Calculated electrostatic bridges between the surface chemical groups from oil and
the surface of the muscovite. Fig. 8. Contact angle vs Z parameter with various oil/brine/rock system.
402 E. Pooryousefy et al. / Journal of Molecular Liquids 250 (2018) 396–403

Table 2
Oil/brine/rock system data from previous work [17,42,45] and this work.

System pH Zeta potential Z* (mV) Contact angle (degree)

Brine Crude Oil Substrate Rock/brine Oil/brine

This work 2000 mg/l CaCl2 Texas oil Mica 7.0 4.94 −2.81 2.13 118
50,000 mg/l CaCl2 Mica 7.3 6.92 14.73 21.65 83
Xie et al. work [42] Formation brine Tarim oil Sandstone 6.2 −3.56 −1.89 5.45 53
Softened brine Sandstone 7.7 −6.60 −3.30 9.90 39
100 diluted formation brine Sandstone 6.1 −19.27 −8.68 27.95 28
Xie et al. work [45] Formation brine Changqing oil Sandstone 7.1 −8.70 −7.70 16.40 74
Ion tuning water Sandstone 8.5 −21.70 −18.70 40.40 39
2000 ppm NaCl Sandstone 8.4 −23.70 −21.20 44.90 37
5800 ppm NaCl Sandstone 7.5 −12.00 −10.20 22.20 54
100,000 ppm NaCl Sandstone 7.2 −8.80 −7.70 16.50 75
Nasralla and Nasr-El-Din work [17] Sea water Crude oil ([17], Table 2) Mica 7.6 2.0 −2.0 0.0 75
5000 mg/l CaCl Mica 5.5 4.0 −5.0 1.0 45
5000 mg/l NaCl Mica 5.9 −12.0 −13.0 25.0 30
10% AQ Mica 7.3 −13.0 −22.0 35.0 27
10% AQ Mica 4.8 −8.0 −10.0 18.0 58

Note: the acid number and base number for Tarim oil was 3.98 and 1.3 mg KOH/g [42]. Yet, the acid number and base number for Changqing oil was not available in the literature [45]. The
acid number and base number for Nasralla and Nasr-El-Din experiments was 0.18 mg KOH/g, and 1.65 mg HCl/g [17].

4. Conclusions Competing financial interests

Understanding of the wettability of oil/brine/rock system is of vital The authors declare no competing financial interests.
importance to subsurface multiphase flow behaviour. To decipher the
factors controlling the wetting characteristics of reservoirs, we mea- Appendix A. Contact angle measurement data were provided below
sured the contact angle of an oil (AN = 1.58 mg KOH/g, BN =
1.02 mg KOH/g) at the surface of muscovite sheet in the presence of
CaCl2 with various salinity levels (2000, 10,000, 30,000 and
50,000 mg/l) at temperatures (50 and 100 °C). We also measured the Salinity, mg/l Contact angle
zeta potential of oil/brines and brines/muscovite. Moreover, we per-
50 °C 100 °C
formed a surface complexation modelling to better understand the fac-
tors governing the wetting characteristics of the oil/brine/muscovite 2000 114 118
10,000 100 102
system. 30,000 86 89
Our results show that the contact angle of oil droplet on the musco- 50,000 78 83
vite surface decreased with increasing salinity of CaCl2, implying that in-
creasing salinity level shifted wettability of muscovite from oil-wet to
less water-wet. Zeta potential results supported the contact angle
References
tests, showing that the polarity of the zeta potential of brine/oil was
shifted from negative to positive with increasing the salinity level [1] J.S. Buckley, N.R. Morrow, Characterization of Crude Oil Wetting Behavior by Adhe-
from 2000 to 50,000 mg/l, but the polarity of the zeta potential of sion Tests, in: SPE/DOE Enhanced Oil Recovery Symposium, 1990 Copyright 1990,
Society of Petroleum Engineers, Inc., Tulsa, Oklahoma, 1990.
brine/muscovite remained positive. This suggests that 2000 mg/l CaCl2
[2] G.J. Hirasaki, Wettability:fundamentals and surface forces, SPE Form. Eval. 6 (1991)
triggered attractive electrical double layer force, yet the 50,000 mg/l 217–226.
CaCl2 rendered repulsive electrical double layer force. Surface complex- [3] G.P. Willhite, in: Richardson (Ed.), Waterflooding, Society of Petroleum Engineers,
TX, 1986.
ation modelling show that at pH = 7, the electrostatic bridges de-
[4] E.W. Al-Shalabi, K. Sepehrnoori, A comprehensive review of low salinity/engineered
creased with the increase of the salinity level, implying the wettability water injections and their applications in sandstone and carbonate rocks, J. Pet. Sci.
of rock surface shifting towards water-wet, consistent with the contact Eng. 139 (2016) 137–161.
angle results. Together, we conclude that DLVO theory is consistent with [5] P.C. Myint, A. Firoozabadi, Thin liquid films in improved oil recovery from low-salin-
ity brine, Curr. Opin. Colloid Interface Sci. 20 (2015) 105–114.
the surface complexation modelling because of the same nature of the [6] J.J. Sheng, Critical review of low-salinity waterflooding, J. Pet. Sci. Eng. 120 (2014)
physics. Therefore, we proposed a new parameter, Z* (Z∗ = | Zetaoil/ 216–224.
[7] G. Jerauld, K. Webb, C.Y. Lin, J. Seccombe, Modeling low-salinity waterflooding, SPE
brine + Zetabrine/rock |) to relate the contact angle with zeta potential.
Reserv. Eval. Eng. 11 (2008) 1000–1012.
We show that the contact angle of oil on the surface of rock decreases [8] A. Lager, K. Webb, C.J. Black, M. Singleton, K. Sorbie, Low salinity oil recovery-an ex-
linearly with increasing Z*. We therefore argue that the manipulation perimental investigation1, Petrophysics 49 (2008).
of Z* (polarity of zeta potential of brine/rock and oil/brine) is of vital im- [9] A.M. Al-sofi, A.A. Yousef, Insight into Smart-water Recovery Mechanism Through
Detailed History Matching of Coreflood Experiments, Society of Petroleum Engi-
portance to shift the reservoir wettability. We suggest that the Z* may neers, 2013.
be used as an interpolant to model the low salinity water flooding, rath- [10] S.C. Ayirala, E. Uehara-Nagamine, A.N. Matzakos, R.W. Chin, P.H. Doe, P.J. van den
er than using salinity level or individual ion in the solution. Hoek, A Designer Water Process for Offshore Low Salinity and Polymer Flooding Ap-
plications, Society of Petroleum Engineers, 2014.
[11] A. Cense, S. Berg, K. Bakker, E. Jansen, Direct Visualization of Designer Water
Acknowledgement Flooding in Model Experiments, Society of Petroleum Engineers, 2014.
[12] G.-Q. Tang, N.R. Morrow, Influence of brine composition and fines migration on
crude oil/brine/rock interactions and oil recovery, J. Pet. Sci. Eng. 24 (1999) 99–111.
This work is supported by Open Fund (PLN201603) of State Key Lab-
[13] P. McGuire, J. Chatham, F. Paskvan, D. Sommer, F. Carini, Low Salinity Oil Recovery:
oratory of Oil and Gas Reservoir Geology and Exploitation (Southwest An Exciting New EOR Opportunity for Alaska's North Slope, 2005.
Petroleum University-China). Support from Sandia National Laborato- [14] S.Y. Lee, K.J. Webb, I. Collins, A. Lager, S. Clarke, M. O'Sullivan, A. Routh, X. Wang, Low
ries is also highly appreciated. We would also like to extend our appre- salinity oil recovery: increasing understanding of the underlying mechanisms, SPE
Improved Oil Recovery Symposium, Tulsa, Oklahoma, USA, 2010.
ciation to Mr. Bob Webb for his help and support towards conducting [15] J. Seccombe, A. Lager, K. Webb, G. Jerauld, E. Fueg, Improving Wateflood Recovery:
the contact angle measurements. LoSalTM EOR Field Evaluation, 2008.
E. Pooryousefy et al. / Journal of Molecular Liquids 250 (2018) 396–403 403

[16] K.J.W.A. Lager, C.J.J. Black, M. Singleton, K.S. Sorbie, Low Salinity Oil Recovery-an Ex- [40] R.A. Nasralla, H.A. Nasr-El-Din, Coreflood study of low salinity water injection in
perimental Investigation, SCA2006-36, 2006. sandstone reservoirs, SPE/DGS Saudi Arabia Section Technical Symposium and Exhi-
[17] R.A. Nasralla, H.A. Nasr-El-Din, Double-layer expansion: is it a primary mechanism bition, Al-Khobar, Saudi Arabia, 2011.
of improved oil recovery by low-salinity waterflooding? SPE Reserv. Eval. Eng. 17 [41] J.J. Sheng, Critical review of low-salinity waterflooding, J. Pet. Sci. Eng. 120 (2014)
(2014) 49–59. 216–224.
[18] A. RezaeiDoust, T. Puntervold, S. Strand, T. Austad, Smart water as wettability mod- [42] Q. Xie, P.V. Brady, E. Pooryousefy, D. Zhou, Y. Liu, A. Saeedi, The low salinity effect at
ifier in carbonate and sandstone: a discussion of similarities/differences in the high temperatures, Fuel 200 (2017) 419–426.
chemical mechanisms, Energy Fuel 23 (2009) 4479–4485. [43] P.V. Brady, N.R. Morrow, A. Fogden, V. Deniz, N. Loahardjo, Winoto, Electrostatics
[19] M. Lashkarbolooki, M. Riazi, F. Hajibagheri, S. Ayatollahi, Low salinity injection into and the low salinity effect in sandstone reservoirs, Energy Fuel 29 (2015) 666–677.
asphaltenic-carbonate oil reservoir, mechanistical study, J. Mol. Liq. 216 (2016) [44] J.M. Schembre, G.Q. Tang, A.R. Kovscek, Wettability alteration and oil recovery by
377–386. water imbibition at elevated temperatures, J. Pet. Sci. Eng. 52 (2006) 131–148.
[20] K. Sandengen, O. Arntzen, Osmosis during low salinity water flooding, IOR 2013- [45] Q. Xie, Y. Liu, J. Wu, Q. Liu, Ions tuning water flooding experiments and interpreta-
from Fundamental Science to Deployment, 2013. tion by thermodynamics of wettability, J. Pet. Sci. Eng. 124 (2014) 350–358.
[21] J.J. Sheng, 3.4.2.3 Multicomponent Ion Exchange, in: Modern Chemical Enhanced Oil [46] A. Mohsenzadeh, P. Pourafshary, Y. Al-Wahaibi, Oil Recovery Enhancement in Car-
Recovery - Theory and Practice, Elsevier, 2011. bonate Reservoirs Via Low Saline Water Flooding in Presence of Low Concentration
[22] E.W. Al-Shalabi, A Comprehensive Review of low salinity/engineered water injec- Active Ions; A Case Study, In; SPE EOR Conference at Oil and Gas West Asia, 2016.
tions and their applications in sandstone and carbonate rocks, J. Pet. Sci. Eng. 139 [47] S. Takahashi, A.R. Kovscek, Wettability estimation of low-permeability, siliceous
(2015) 137–161. shale using surface forces, J. Pet. Sci. Eng. 75 (2010) 33–43.
[23] J. Wu, F.H. Liu, G. Chen, X. Wu, D. Ma, Q. Liu, S. Xu, S. Huang, T. Chen, W. Zhang, The [48] G.J. Hirasaki, Wettability:fundamentals and surface forces, SPE Form. Eval. 6 (1991)
effect of ionic strength on the interfacial forces between oil/brine/rock interfaces: a 217–226.
chemical force microscopy study, Energy Fuel 30 (2015) 273–280. [49] Q. Xie, A. Saeedi, E. Pooryousefy, Y. Liu, Extended DLVO-based estimates of surface
[24] J. Matthiesen, N. Bovet, E. Hilner, M.P. Andersson, D.A. Schmidt, K.J. Webb, K.N. force in low salinity water flooding, J. Mol. Liq. 221 (2016) 658–665.
Dalby, T. Hassenkam, J. Crouch, I.R. Collins, S.L.S. Stipp, How naturally adsorbed ma- [50] C. Busireddy, D.N. Rao, Application of DLVO theory to characterize spreading in
terial on minerals affects low salinity enhanced oil recovery, Energy Fuel 28 (2014) crude oil-brine-rock systems, SPE/DOE Symposium on Improved Oil Recovery, Soci-
4849–4858. ety of Petroleum Engineers, Tulsa, Oklahoma, 2004.
[25] G. Yang, T. Chen, J. Zhao, D. Yu, F. Liu, D. Wang, M. Fan, W. Chen, J. Zhang, H. Yang, J. [51] A. Lager, K.J. Webb, C.J.J. Black, M. Singleton, K.S. Sorbie, Low salinity oil recovery - an
Wang, Desorption mechanism of Asphaltenes in the presence of electrolyte and the experimental investigation, Petrophysics 49 (2008) 28–35.
extended Derjaguin–landau–Verwey–Overbeek theory, Energy Fuel 29 (2015) [52] G. Sposito, The Chemistry of Soils, Oxford University Press, 1989.
4272–4280. [53] P.V. Brady, J.L. Krumhansl, Surface complexation modeling for waterflooding of
[26] H. Mahani, A.L. Keya, S. Berg, R. Nasralla, Electrokinetics of Carbonate/Brine Interface sandstones, SPE J. 18 (2013).
in Low-Salinity Waterflooding: Effect of Brine Salinity, Composition, Rock Type, and [54] Z. Aghaeifar, S. Strand, T. Austad, T. Puntervold, H. Aksulu, K. Navratil, S. Storås, D.
pH on ζ-Potential and a Surface-Complexation Model, 2016. Håmsø, Influence of formation water salinity/composition on the low-salinity en-
[27] P.V. Brady, G. Thyne, Functional wettability in carbonate reservoirs, Energy Fuel 30 hanced oil recovery effect in high-temperature sandstone reservoirs, Energy Fuel
(2016) 9217–9225. 29 (2015) 4747–4754.
[28] T. Hassenkam, J. Mathiesen, C. Pedersen, K. Dalby, S. Stipp, I.R. Collins, Observation of [55] P.V. Brady, C.R. Bryan, G. Thyne, H. Li, Altering wettability to recover more oil from
the low salinity effect by atomic force adhesion mapping on reservoir sandstones, tight formations, J. Unconv. Oil Gas Resour. 15 (2016) 79–83.
SPE Improved Oil Recovery Symposium, Society of Petroleum Engineers, Tulsa, [56] H. Hartridge, R.A. Peters, Interfacial tension and hydrogen-ion concentration, Proc. R.
Oklahoma, USA, 2012. Soc. London, Ser. A 101 (1922) 348–367.
[29] D. Kuznetsov, S. Cotterill, M.A. Giddins, M.J. Blunt, Low-Salinity Waterflood Simula- [57] J.F. Danielli, The relations between surface pH, ion concentrations and interfacial
tion: Mechanistic and Phenomenological Models, Society of Petroleum Engineers, tension, Proc. R. Soc. London, Ser. B 122 (1937) 155–174.
2015. [58] N.J. Hadia, T. Hansen, M.T. Tweheyo, O. Torsæter, Influence of crude oil components
[30] H. Mohammadi, G. Jerauld, Mechanistic modeling of the benefit of combining poly- on recovery by high and low salinity waterflooding, Energy Fuel 26 (2012)
mer with low salinity water for enhanced oil recovery, SPE Improved Oil Recovery 4328–4335.
Symposium, Society of Petroleum Engineers, Tulsa, Oklahoma, USA, 2012. [59] J.N. Israelachvili, Chapter 14 - Electrostatic forces between surfaces in liquids, in: J.N.
[31] C. Qiao, R. Johns, L. Li, Modeling low-salinity waterflooding in chalk and limestone Israelachvili (Ed.), Intermolecular and Surface Forces, Third EditionAcademic Press,
reservoirs, Energy Fuel 30 (2016) 884–895. San Diego 2011, pp. 291–340.
[32] C. Dang, L. Nghiem, N. Nguyen, Z. Chen, Q. Nguyen, Mechanistic modeling of low sa- [60] R.A. Nasralla, H.A. Nasr-El-Din, Impact of cation type and concentration in injected
linity water flooding, J. Pet. Sci. Eng. 146 (2016) 191–209. brine on oil recovery in sandstone reservoirs, J. Pet. Sci. Eng. 122 (2014) 384–395.
[33] C.V. Guidotti, Micas in metamorphic rocks, Rev. Mineral. Geochem. 13 (1984) [61] H. Mahani, A.L. Keya, S. Berg, W.-B. Bartels, R. Nasralla, W.R. Rossen, Insights into the
357–467. mechanism of wettability alteration by low-salinity flooding (LSF) in carbonates,
[34] G. Hansen, A.A. Hamouda, R. Denoyel, The effect of pressure on contact angles and Energy Fsuel 29 (2015) 1352–1367.
wettability in the mica/water/n-decane system and the calcite + stearic acid/ [62] Q. Xie, D. Ma, Q. Liu, J. Wu, Primary mechanism of improved oil recovery by low sa-
water/n-decane system, Colloids Surf. A Physicochem. Eng. Asp. 172 (2000) 7–16. linity waterflooding: double layer expansion or multicomponent ionic exchange?
[35] T. Austad, A. Rezaeidoust, T. Puntervold, Chemical Mechanism of Low Salinity Water in: S.O.C. Analysts (Ed.), SCA2013-85, 2013 Napa California.
Flooding in Sandstone Reservoirs, Society of Petroleum Engineers, 2010. [63] Q. Xie, J. Wu, J. Qin, Q. Liu, D. Ma, Investigation of electrical surface charges and wet-
[36] T. Austad, A. Rezaeidoust, T. Puntervold, Chemical mechanism of low salinity water tability alteration by ions matching waterflooding, International Symposium of the
flooding in sandstone reservoirs, SPE Improved Oil Recovery Symposium, 24–28 Society of Core Analysts, 2012.
April, Tulsa, Oklahoma, USA, 2010. [64] M.D. Jackson, D. Al-Mahrouqi, J. Vinogradov, Zeta potential in oil-water-carbonate
[37] A. RezaeiDoust, T. Puntervold, T. Austad, Chemical verification of the EOR mecha- systems and its impact on oil recovery during controlled salinity water-flooding,
nism by using low saline/smart water in sandstone, Energy Fuel 25 (2011) Sci. Rep. 6 (2016) 37363.
2151–2162. [65] H. Mahani, A.L. Keya, S. Berg, R. Nasralla, Electrokinetics of carbonate/brine Interface
[38] M.E. Haagh, I. Siretanu, M.H. Duits, F. Mugele, Salinity-dependent contact angle al- in low-salinity waterflooding: effect of brine salinity, composition, rock type, and pH
teration in oil/brine/silicate systems: the critical role of divalent cations, Langmuir on ζ-potential and a surface-complexation model, SPE J. 22 (2016) 53–68.
33 (2017) 3349. [66] Q. Xie, Y. Chen, A. Sari, W. Pu, A. Saeedi, X. Liao, A pH-resolved wettability alteration:
[39] M. Frieder, B. Bijoyendra, C. Andrea, S. Igor, M. Armando, D. Michel, C.-S. Martien, E. implications for CO2-assisted EOR in carbonate reservoirs, Energy Fuel (2017).
Dirk Van Den, S. Isabella, C. Ian, Ion adsorption-induced wetting transition in oil-
water-mineral systems, Sci. Rep. 5 (2015).

You might also like