You are on page 1of 20

Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Adsorption analysis of natural anionic surfactant for enhanced oil recovery: T


The role of mineralogy, salinity, alkalinity and nanoparticles
Neha Saxena, Amit Kumar, Ajay Mandal∗
Department of Petroleum Engineering, Indian Institute of Technology (Indian School of Mines), Dhanbad, 826 004, India

A R T I C LE I N FO A B S T R A C T

Keywords: Anionic surfactants are widely used as an effective chemical for enhanced oil recovery because of their unique
Anionic surfactant characteristics of reducing interfacial tension (IFT) between the trapped crude oil and water, and alteration of
Reservoir rock wettability of reservoir rock from oil-wet to water-wet. However, the loss of surfactant by adsorption onto the
Adsorption solid rock surface is a major concern that reduces the efficiency of the surfactant flooding process and must be
Isotherm models
considered while designing the process. Present study highlights the equilibrium adsorption and kinetics of an
Nanoparticles
anionic surfactant synthesized from soap-nut fruit on sandstone, carbonate and bentonite clay as representative
of reservoir rocks. The mineralogy and morphology of the rocks were investigated by FE-SEM, XRD and BET
surface area analysis. UV–Visible spectroscopy was used to measure the amount of surfactant adsorbed on solid
rock/clay surface in batch experiments. The experimental equilibrium adsorption data were analyzed by using
Langmuir, Freundlich, Temkin and Redlich-Peterson isotherm models and adsorption parameters were calcu-
lated. Adsorption kinetics of the surfactant system were studied and pseudo second order kinetic model was best
fitted for the surfactant/rock/clay systems. Presence of salt was found to increase surfactant adsorption on rock/
clay marginally. However, presence of alkali and nanoparticles was found to reduce the loss of surfactant by
adsorption and shows synergistic effects on IFT reduction, which is beneficial for application of the surfactant in
oil recovery. The findings of the study are quite helpful in proper designing of the surfactant flooding for en-
hanced oil recovery.

1. Introduction which prevents the contact of crude oil and rock surface (Hiorth et al.,
2010). The rupture of the brine layer occurs due to the presence of
Surfactants are amphiphilic compounds that have a hydrophilic attractive forces such as charge transfer, van der Waals forces and hy-
head with a hydrophobic tail. The amphiphilic structure of the surfac- drogen bonding between crude oil and rock surface (Hirasaki, 1991).
tants has led to their unique properties which have led to its use in Thus, these forces lead to adsorption of crude oil components on the
various industrial application like detergents, foaming agents, wetting rock surface and oil-wetting state of the reservoir rock. The state of
agents, emulsifiers, and oil industry. The application of surfactants in reservoir rock surface before surfactant flooding can be seen in the
oil industry is mostly driven by its property to reduce IFT of oil-water schematic diagram depicted in Fig. 1. When surfactant is injected in the
interface and with the adsorption of surfactant molecules on oil wet reservoir, the surfactant molecules interact with the adsorbed compo-
surface, it leads to alteration of wettability of rock. These phenomena nents of the crude oil like naphthenic acid, asphaltenes and resins.
collectively lead to the recovery of trapped oil from the reservoir. These adsorbed components of crude oil, which were responsible for oil
The wetting nature of a reservoir is usually intermediate to oil wet wetting state of surface, can have either hydrophilic interaction or
and it is dependent on the rock mineralogy, composition of crude oil hydrophobic interaction with the injected surfactant molecules. Hy-
and brine, and temperature and pressure conditions of reservoir (Sohal drophilic interaction leads to formation of ion pair between the hy-
et al., 2016). The mechanism of wettability alteration of originally drophilic head of the surfactant and adsorbed crude oil (Hou et al.,
water wet reservoir to oil wetting have been related to adhesion of 2015). The bonding of the carbon chain of surfactant molecules with
polar components of crude oil, which had migrated into the reservoir the adsorbed crude oil components leads to hydrophobic interaction
(Rahbar et al., 2012). After the migration of crude oil into reservoir between crude oil and surfactant molecules (Standnes and Austad,
pores, there exists a layer of brine between crude oil and rock surface, 2000). With the movement of the flood front, the crude oil components


Corresponding author.
E-mail address: ajay@iitism.ac.in (A. Mandal).

https://doi.org/10.1016/j.petrol.2018.11.002
Received 5 September 2018; Received in revised form 1 November 2018; Accepted 1 November 2018
Available online 02 November 2018
0920-4105/ © 2018 Elsevier B.V. All rights reserved.
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 1. (a) Schematic of preferential adsorption of surfactant on oil aged sandstone rock with wettability alteration; (b) Change of Zeta potential on adsorption of
surfactant on oil aged sandstone rock.

bonded with the surfactant molecules are desorbed into the bulk of the decades have been carried out to analyze the adsorption behavior of
aqueous phase and the available site gets occupied by surfactant mo- various ionic and non-ionic surfactant onto the rock surface. The phy-
lecule (Fig. 1a). This can also be seen in the schematic diagram where at siochemical properties that affect the adsorption of surfactant are pH,
low surfactant concentration, there is partial replacement of adsorbed temperature, ionic strength and electrolyte concentration in the re-
crude oil components with surfactant molecules. With increase in sur- servoir. Any variation in the mentioned physiochemical properties can
factant concentration, most of the adsorbed crude oil are desorbed and result in change in adsorption pattern of the system. It is widely known
trapped in the surfactant micelles. The adsorption of surfactant mole- that the adsorption of ionic surfactants onto the rock surface is gov-
cules leads to increase in zeta potential of surface as shown in Fig. 1b erned by electrostatic attraction. The similarity of charges of anionic
(Saxena et al., 2018a). This increase in the surface charge is due to the surfactant and sandstone reservoir surface leads to preference of an-
dominance of negative charge by sulfonate ions and hydroxyl groups. ionic surfactants over nonionic and cationic surfactants for applicability
Thus, surfactant molecules adsorbed on the rock surface prevent further in sandstone reservoirs. Scientists throughout the world have worked in
interaction of crude oil and rock surface and changes the wettability of last few years to reduce the surfactant adsorption on to the rock surface
rock to water wet. (Bournival et al., 2104; Yekeen et al., 2016; Saha et al., 2017; Bera
However, excess adsorption of surfactant leads to reduction in their et al., 2013; Bai and Grigg, 2005; Somasundaran and Hanna, 1979).
concentration and limits their efficiency during the surfactant based Traditionally sodium carbonate and sodium hydroxide as alkaline re-
EOR process. Thus, a comprehensive knowledge of loss of surfactant by agents are used to maintain the negative charge of the rock surface
adsorption on the reservoir rock is of utmost importance prior to the causing electrostatic repulsion between rock surface and the surfactant
injection of surfactant slugs for chemical EOR. Extensive studies in past molecules leading to significant decrease in surfactant adsorption in

1265
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 2. Reaction scheme for the synthesis of Soap-nut surfactant.

Fig. 3. Absorbance of the surfactant solution at varying surfactant concentration.

sandstone reservoirs (Gregersen et al., 2013). Gogoi (Gogoi, 2009) found by some researchers that additives like alkali and polymer are
studied the effect of salt concentration and pH on adsorption of sur- ineffective in reducing the adsorption in high-saline and high tem-
factant on the reservoir rock and reported that with the increase in salt perature reservoirs (Al-hashim et al., 1996; Gupta and Mohanty, 2007;
concentration, adsorption of surfactant increases and increase in pH Wang et al., 2013). Henceforth, the prevention of adsorption of sur-
leads to decrease in the surfactant adsorption. Wang et al. (Wang et al., factant onto the reservoir rock surface is a challenging task and it is
2015) showed that presence of polymers also lowers the adsorption of important to make it an economically viable process for its application
surfactant by forming a thick protective layer over the rock surface and in oil recovery from reservoir.
reduces the active sites available for adsorption. However, it has been The main objective of the present paper is to study the adsorption

1266
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 4. Calibration curve of surfactant solution at wavelength 259 nm.

Fig. 5. Calibration curve for the surfactant solution using conductivity for sandstone rock surface.

Table 1
Properties of the adsorbent used in the adsorption study.
Adsorbent BET SurfaceArea (m2/g) t-plot Micro-pore Area (m2/g) t-plot External Surface Area (m2/g) t-plot Micro-pore Volume (cm3/g) Average Pore Diameter Å

−5
Sandstone 0.51 0.144 0.366 7.5 × 10 19.96
Carbonate 2.80 1.35 1.43 6.9 × 10−4 21.61
Bentonite 51.72 10.71 41.03 5.05 × 10−3 20.95

pattern of anionic surfactant on to rock surface. The adsorption study 2. Materials required
was conducted using sandstone, carbonate as a prototype of the re-
servoir rock systems. Bentonite was chosen as it is a clay containing The surfactant used in this study is a natural surfactant, synthesized
montmorillonite and they are present in large amount in typical shaly from soap-nut oil through trans-esterification followed by sulfonation
clastic reservoir. The adsorbate samples were characterized by Field process. The synthesis process has been discussed in our earlier paper
emission scanning electron microscopy (FE-SEM), X-Ray Diffraction (Saxena et al., 2018b). The chemical structure along with the synthesis
(XRD) and gas adsorption analysis to measure the surface area available scheme is shown in Fig. 2. The chemical structure of the synthesized
for adsorption. Anionic surfactant synthesized from soap-nut fruit was surfactant (compound 2) was confirmed by GC-MS and FTIR analysis.
selected as adsorbate in the presented study. Effects of salinity, rock The particular surfactant was chosen for the study as the surfactant had
type, alkali and nanoparticles were studied to analyze their effect in shown ultralow IFT and it is also cost effective and biodegradable in
altering the adsorption of surfactant on the rock surface for making it an nature. The surfactant solution was also found to be thermally stable up
economical system for its application in the field of EOR. to reservoir temperature.
Sodium chloride and sodium carbonate were obtained from SRL to

1267
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 6. X-Ray Diffractogram images of rock (a) Sandstone, (b) Carbonate and (c) Bentonite.

study the effect of salinity and alkalinity on surfactant adsorption. Silica sieved to 60–80 mesh. The bentonite clay used in the experiments was
nanoparticle of size 15–20 nm and alumina nanoparticles of size purchased from the market and was used in its original form. These
20–30 nm procured from SRL were used to study the effect of nano- powders were not washed as it may dissolve the existing components of
particles on adsorption phenomenon. Sandstone and carbonate rock the rock sample. The powdered samples were used for characterization
samples were collected from Cambay Basin, India. Bentonite clay used and adsorption study. FE-SEM and EDX analysis (using SUPRA -55
in the experiment was obtained from Loba Chemicals Pvt. Ltd. The ZEISS, Germany) of the rock sample were performed to study the
crude oil used in the experiments was procured from Oil India Limited, morphology and the elemental composition of the adsorbent used in the
Assam. The obtained crude oil has a total acid number 0.84 mg KOH/g, analysis. Brunauer–Emmett–Teller (BET) analysis (using Micrometrics
gravity of 18.9° API, density of 0.924 g/cc and viscosity of 4.26 cP at ASAP, 2020) was done to calculate the surface area available for ad-
27 °C. The SARA composition of the crude oil sample indicated 19.6% sorption using liquid nitrogen at degassing temperature of 150 °C. X-
of saturates, 77% of aromatics, and 3.1% of resin and asphaltenes. Ray diffractogram of crushed rock samples were analyzed by measuring
the Bragg angle 2θ (10°≤ 2θ ≤ 90°) using XRD instrument (Bruker D8
3. Experimental advanced) with Cu as target radiation of wavelength (λ = 0.15405 nm).
The zeta potential of the suspension of the powdered samples was
3.1. Characterization of adsorbent samples tested by Horiba Nano Particle Analyser SZ-100. The pH of the sus-
pension was maintained to acidic range using 0.1N HCl and basic pH
Rock samples of sandstone and carbonate rock were grinded and using 0.1N KOH.

1268
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 7. FE-SEM images (a) Sandstone, (b) Carbonate and (c) Bentonite at 10.00 KX magnification.

liquid was estimated by using the calibration curve, using the surfactant
solutions of known concentrations. The initial and final concentration
of the surfactant solutions were used to calculate the adsorbed amount
of adsorbent at equilibrium concentrations as per Eq. (1):
V (Ci − Ce )
q=
m (1)

where, Ci and Ce are the initial and final equilibrium concentration of


the surfactant in the solution (mg/L) respectively. V represents the
volume of the surfactant solution (L) and mass of the adsorbent (g) used
in the experimental analysis is represented by m. Fig. 3 depicts the
maximum absorbance obtained at the wavelength of 259 nm for the
surfactant solutions of varying concentrations. The maximum absor-
bance obtained for the respective surfactant concentration was used to
plot the calibration curve as shown in Fig. 4. The data obtained from
the adsorption study was fitted to various static and kinetic models and
the respective parameters of the models was used to study the ad-
sorption behavior of surfactant.
To ensure that the adsorption study using UV–Visible spectroscopy
Fig. 8. Zeta potential of rock surface at varying pH. shows consistency in results, the adsorption was also confirmed by
conductivity. The conductivity meter (Hanna HI-2003) was used to
3.2. Adsorption studies measure the conductance of the surfactant solution. These conductance
values were used to plot the calibration curve for the standard surfac-
The grinded rock and clay samples were added to the surfactant tant solution at different concentration from 1000 mg/L to 15000 mg/L.
solutions of concentration ranging 1000–15000 mg/L. The solutions This technique is also used by other research group around the world to
were subjected to rotation at 50 rpm for 24 h using Rotospin (Tarsons). study adsorption behavior of the surfactant onto the rock surface
After 24 h the samples were segregated using centrifuge at 3000 rpm for (Barati-Harooni et al., 2016; Ahmadi and Shadizadeh, 2013; Barati
20 min. Liquid obtained at the upper portion was decanted after cen- et al., 2016). Experiments using conductivity method were conducted
trifugation and was analyzed under UV–Visible spectrophotometer and compared with the results obtained from the UV–Visible spectro-
(SHIMADZU UV-1800) to obtain the equilibrium concentration. The UV scopy. The calibration curve for the surfactant solution on the sand-
spectroscopy technique is correlated with the interaction of light with stone surface is shown in Fig. 5. The calibration curve obtained by
matter. When a monochromatic light is absorbed by compound, it leads conductivity method was then used to study the adsorption on sand-
to increase of energy content of the atomic or molecular constituents of stone powder and compared with the result obtained by UV spectro-
the chemical compound. Thus, for every compound, a peak is obtained scopy method. However, the conductivity method cannot be used to
at specific wavelength. The unknown concentration of the supernatant test the adsorption in presence of salt, alkali and nanoparticle as pre-
sence of these additional constituents changes the values of

1269
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 9. Comparison of adsorption of surfactant on sandstone rock surface using two different techniques.

Fig. 10. Equilibrium profiles for adsorption of soap-nut surfactant against the initial concentration.

Fig. 11. Variation in IFT at oil-aqueous interface after adsorption.

1270
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 12. Langmuir adsorption isotherm fitting curves of (a) Sandstone (b) Carbonate (c) Bentonite.

1271
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 13. Freundlich adsorption isotherm fitting curves of (a) Sandstone (b) Carbonate (c) Bentonite.

1272
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 14. Temkin adsorption isotherm fitting curves of (a) Sandstone (b) Carbonate (c) Bentonite.

conductance. Thus, the conductance values of supernatant surfactant 3.3. IFT study of soap-nut surfactant before and after adsorption
solution will not correspond to the actual surfactant concentration of
the solution. The measurement of IFT before and after adsorption provides an
indication of loss of IFT reduction capability of surfactant which occurs
due to loss of surfactant. The IFT between surfactant solution and crude

1273
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 15. Redlich- Peterson adsorption isotherm fitting curves of (a) Sandstone (b) Carbonate (c) Bentonite.

oil was measured using spinning drop tensiometer (SVT15 Data phy- The process of treating the surfactant solution with rock samples was
sics). The IFT of oil-water interface was measured before adsorption repeated at least 4 times or till IFT in ultralow range was obtained.
and compared with the IFT obtained for the supernatant surfactant Greater repetition of process signifies greater tolerance of surfactant to
solution collected after the adsorption study. The supernatant surfac- adsorption loss.
tant liquid was retreated with fresh adsorbent samples for 12 h and the
surfactant solution was recollected to check its IFT reduction capability.

1274
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Table 2
Adsorption parameters of isotherm model of adsorption of soap-nut surfactant at rock interface.
Isotherm Model Parameters Sandstone Carbonate Bentonite

2
Langmuir Model R 0.9954 0.9677 0.9998
KL(L/mg) 6.6 × 10−4 1.59 × 10−3 9.24 × 10−4
qo(mg/g) 45.45 50.12 67.56
Freundlich Model R2 0.9601 0.9689 0.9324
b(mg/g) 0.4932 0.952 0.9514
n 2.04 2.12 2.07
Temkin Model R2 0.9811 0.9453 0.9956
B 10.11 12.40 13.42
Km(L/g) 6.53 × 10−3 1.1 × 10−2 1.1 × 10−2
Redlich-PetersonModel R2 0.9957 0.9825 0.9992
K (L/mg) 3.5 × 10−1 2.08 × 10−1 6.01 × 10−2
α (L/mg)β 3.7 × 10−3 1.1 × 10−1 9.02 × 10−4
β 0.819 0.607 1

7000 mg/L
60
(a) Sandstone Carbonate bentonite
Amount adsorbed mg/g

50

40

30

20

10

0
0% 1% 2% 3%
Salt content (%)

9000 mg/L
70
Sandstone Carbonate bentonite
60 (b)
Amount adsorbed mg/g

50

40

30

20

10

0
0% 1% 2% 3%
Salt content (%)

Fig. 16. Effect of salt content on adsorption of surfactant molecules at concentration of (a) 7000 mg/L (b) 9000 mg/L.

1275
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 17. Effect of alkali on adsorption of surfactant molecules at concentration (a) 7000 mg/L (b) 9000 mg/L.

4. Result and discussion 4.1.2. XRD analysis


Adsorbent sample usually possesses a charge on their surface oc-
4.1. Characterization of rocks/clay curring due to the mineralogy and crystalline-amorphous structure of
the adsorbent. The presence of these surface charges affects the cou-
4.1.1. BET analysis lombic forces between the adsorbate molecules and adsorbent surface.
The amount of adsorption of an adsorbate on adsorbent is directly Thus, XRD analysis was used for the identification of mineralogical
correlated to the available surface area of the adsorbent. BET analysis constituents (e.g. minerals, or inorganic compounds) of adsorbent
measures the dissolution rate, which is proportional to the specific sample. Fig. 6 (a) shows the X-ray diffraction pattern of the crushed
surface area. Thus, this surface area helps to predict the availability of sandstone sample. Characteristic peak for the quartz are obtained at
porous media for adsorption. Table 1 depicts the characteristics of the 21°, 27.74°, 28.57°, 47.23°, 60° with the sharpest peak at 27.47°. All the
rock samples used as an adsorbent for the adsorption study. The BET major peaks represented in the figure corresponds to quartz, thus in-
surface area and average pore volume was calculated by gas adsorption dicating high quartz content with fewer impurities like siderite and
analyser. It was assumed that for the micro-porous adsorbents, the BET kaolinite clay. XRD pattern of carbonate sample presented in Fig. 6 (b)
surface area is the equivalent area rather than actual surface area (Sing shows 2θ peaks of 23.04°, 31.4°, 39.2°, 47.4° and 48.7° which corre-
and Everett, 1973). As depicted in Table 1, the BET area was observed sponds to the presence of calcite and aragonite. Carbonate sample also
to be maximum for bentonite sample and minimum for sandstone contained slight impurities such as quartz for which a lower intensity
sample. The average pore diameter for all the three rock samples under peak was obtained at 27.6°. XRD patterns of fresh bentonite is shown in
considerations was found to be similar. Fig. 6 (c). It is shown in the diffractogram that the 2θ peaks of mon-
tmorillonite are visible at 6.93°, 21.21°, 29.04°, 36.31°, and 66.28°,
whereas 2θ peaks of quartz are identified at 27.74°, 47.23°, 60°. This

1276
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 18. Effect of alumina nanoparticles on adsorption pattern of surfactant molecules at concentration (a) 7000 mg/L (b) 9000 mg/L.

confirms that bentonite clay is a composite mixture with montmor- between the rock surfaces and charged molecules in aqueous medium
illonite and quartz as its major constituents. (Alroudhan et al., 2016). Zeta potential for the sandstone, carbonate
and bentonite clay surface as a function of pH is presented in Fig. 8. It
was observed that for pH variation from 2.3 to 11.5, the rock surface
4.1.3. FE-SEM analysis
showed negative zeta potential indicating that the ionic interaction
FE-SEM analysis of adsorbent sample helps in the study of surface
mechanism for all three surfaces will be same (Hirasaki et al., 2008).
characteristics, such as morphology, texture and roughness. These
The negative charge of sandstone surface is due to dispersion of the
characteristics affect the adsorption capacity of the adsorbent surface
water molecules at the surface and formation of SiOH groups
and film thickness of the adsorbate particle. Thus, the surface of sam-
(Michalske and Freiman, 1983). Zeta potential of natural calcite is
ples was characterized using FE-SEM analysis and captured micro-
controversial as some researchers have reported it as negative, whereas
photographs of fresh sandstone, carbonate and bentonite rock surfaces
some have reported it as positive. This may be due to variation in
are depicted in Fig. 7. The images of sandstone and carbonate depicts
dissociation of calcium ions in water and presence of some amount of
that the sample surfaces have less porous structures as compared to the
quartz in the original composition (Douglas and Walker, 1950; Huang
surface of bentonite. The surface morphology of the samples as pre-
et al., 1991; Monfared et al., 2015). Bentonite sample possesses nega-
dicted by FE-SEM analysis was also supported by the BET results which
tive zeta potential due to the presence of surface hydroxyl groups on the
showed that the bentonite sample had the largest surface area in
basal planes of layered bentonite structure (Baik and Lee, 2010).
comparison to sandstone and carbonate samples.

4.1.4. Effect of pH on charge distribution at rock surface 4.2. Adsorption behavior of surfactant on to the rock
The zeta potential of the natural rock surfaces plays an important
role in various subsurface processes, affecting the ionic interaction The adsorption profile as obtained for the sandstone sample by UV

1277
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 19. Effect of silica nanoparticles on adsorption pattern of surfactant molecules at concentration (a) 7000 mg/L and (b) 9000 mg/L.

spectroscopy and conductivity technique is shown in Fig. 9. It was of plot is steep for surfactant concentration up to its CMC (8000 mg/L),
observed that the two techniques showed comparable results. As stated but for surfactant concentration higher than the CMC, there is a re-
earlier, that the conductivity technique cannot be used for further duction in the slope of the curve.
analysis in presence of salt, alkali and nanoparticle, thus the data ob-
tained by UV spectroscopy technique was used to calculate the ad-
sorption density. The adsorption density as calculated by Eq. (1) was 4.3. IFT study of soap-nut surfactant before and after adsorption
plotted against their initial concentration before adsorption and is
shown in Fig. 10. According to the experimental data, as the surfactant The reduction of IFT of oil-water interface is maximum when the
concentration was increased, more amount of surfactant was adsorbed surfactant concentration is greater than its CMC. However, due to loss
on to the sample surface. At low surfactant concentration, the molecules of surfactant due to adsorption, its concentration in solution decreases
of surfactant are attracted to the charge dispersed on the rock surface as below its CMC, leading to increase in IFT and ineffective displacement
electrical double layer (as explained by Helmontz, 1879; and further of trapped crude oil during flooding process. Thus, IFT between sur-
modified by Stern in 1924) (Helmholtz, 1853; Stern, 1924). This dis- factant solution and crude oil before and after repeated adsorption was
persed charge is responsible for attracting the surfactant molecules. At compared to test for the IFT reduction capability of surfactant against
higher surfactant concentration, the molecules of surfactant start ag- repeated adsorption. The IFT obtained between surfactant solution and
gregating and forming hemi-micelles. When the CMC of the surfactant crude oil have been shown in Fig. 11. As shown in the figure, the IFT of
is reached, the surfactant molecules are present in the bulk as micelles. surfactant solution and crude oil is 2.13 × 10−3 mN/m before adsorp-
The adsorption of these surfactant micelles causes a subsequent in- tion process. The IFT was found to be in ultralow range for 36 h for
crease in the adsorption level at CMC. For surfactant concentration sandstone and carbonate samples and for 24 h for bentonite sample.
higher than CMC, most active sites are adsorbed with surfactant and This signifies that the process of adsorption was repeated 3 times for
micelles of surfactants are repelled by the adsorbed surfactant mole- sandstone and carbonate samples and 2 times for bentonite sample. The
cules. This phenomenon can also be seen from Fig. 10, where the slope greater number of times for which an ultralow IFT in range of
10−2 mN/m is achieved, higher is the resistance of surfactant molecules

1278
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

against adsorption on to the rock (Yan et al., 2017). This indicates that
the surfactant has excellent resistance to adsorption on sandstone and

Weber and Morriss


Wang et al. (2007)
carbonate samples. The IFT reduction property was reduced after ad-
Lagergren (1898) sorption process was repeated third time on bentonite sample due to its
larger surface area leading to greater adsorption as compared to sand-
References

stone and carbonate samples.

(1963)
4.4. Static modelling of adsorption isotherms
Pseudo-first order kinetic equation pioneered by Lagergren in the year 1898 to evaluate the kinetics of liquid and solid phase

The intraparticle mass transfer diffusion model was given in 1963 by Weber and Morris to study the mechanism of diffusion and
In this model, the rate-limiting step is the surface adsorption that involves chemisorption, where the removal from a solution is

The adsorption profile as shown in Fig. 10 was modelled using


Langmuir, Freundlich, Temkin and Redlich-Peterson models and the
parameters of the respective models are analyzed in the following
sections.

4.4.1. Langmuir adsorption isotherm


Langmuir isotherm model was proposed by Irving Langmuir in 1916
which was originally proposed for solid-gas phase bio-sorbents. In its
derivation, Langmuir adsorption isotherm is a homogenous adsorption
where individual molecule has constant enthalpy and activation en-
ergy. This model assumes that the surfactant molecules adsorbs on the
rock surface are limited and there exist no interaction in between the
adsorbed molecules (Langmuir, 1916). The experimental data acquired
from adsorption study of the soap-nut surfactant at varying con-
centration was plotted and was fitted according to the Langmuir
equation represented as Eq. (2). Fig. 12 depicted the Langmuir fitting to
the experimental data at different rock surfaces.
due to physicochemical interactions between the two phases

q o KCe
qe =
KL Ce + 1 (2)
where, qe is amount of adsorbate adsorbed (mg/g); qo is the maximum
amount of adsorbate adsorbed (mg/g); KL is the Langmuir constant (L/
mg); Ce is equilibrium concentration (mg/L). In its linearized form Eq.
(2) can be written as Eq. (3):
1 1 1
= +
qe q0 KL Ce q0 (3)
adsorption process.

4.4.2. Freundlich adsorption isotherm


Fruendlich isotherm was the first sorption model proposed by
adsorption.

Fruendlich in 1906. It was the earliest model proposing a relationship


Comments

between non-ideal and reversible sorption process. The amount ad-


sorbed it the summation of adsorption on all sites each having a specific
bond energy. This model assumes that the adsorbent surface is het-
adsorbed

adsorbed

erogeneous and the adsorption occurs in multiple layers (Freundlich,


k1 (hr−1), rate constant

kid (mg/g/hr), rate


k2 (g/mg/hr), rate

1906). Experimental data obtained by plotting the adsorption curve


Kinetic Parameters

between the amount adsorbed at the rock surface and the equilibrium
concentration of the surfactant solution were fitted using Freundlich
(mg/g),

(mg/g),

model as represented by Eq. (4):


constant

constant
amount

amount

q e = bCe1/n (4)
Various kinetic models to study the rate of adsorption.

qe

qe

where Ce is equilibrium concentration (mg/L), q e is the amount ad-


ln (qe − qt ) = ln(qe ) − k1 t

sorbed onto the rock surface (mg/g). 1/n and b are the Freundlich
adsorption parameters and their values are related with adsorption
capacity and adsorption intensity. Eq. (4) can be written in its simpli-
qe
t

fied linear form as Eq. (5):


+

qt = Kid + C
Equations

k 2 qe2

1
1

log qe = logb + logCe


n (5)
=
qt
t

Lower value of Freundlich constant 0.1 < 1/n < 0.5 predicts that
Intra particle Diffusion

adsorption is favorable and value greater than 0.5 predicts difficulty in


Pseudo Second-Order

adsorption. The fitting of adsorption data to Freundlich model is de-


Pseudo First-Order

picted in Fig. 13.

4.4.3. Temkin adsorption isotherm


Kinetic
Table 3

model

Temkim isotherm model was proposed by Tempkin and Pyzhev in


1940 which explains the effect of interaction between the adsorbate

1279
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Table 4
Kinetics adsorption parameters of surfactant on different rock surface.
Adsorbent Sandstone Carbonate Bentonite

Adsorbate concentration (mg/L) 7000 9000 7000 9000 7000 9000

−1 −2 −2 −2 −2 −2
Model Pseudo First Order k1 (hr ) 6.09 × 10 6.55 × 10 6.24 × 10 7.54 × 10 8.15 × 10 6.57 × 10−2
Qe (mg/g) 60.08 54.40 46.34 71.87 56.33 60.53
R2 0.963 0.987 0.979 0.9164 0.987 0.986
Pseudo Second Order k2 (g/mg/hr) 5.03 × 10−5 5.54 × 10−5 4.93 × 10−5 2.92 × 10−5 4.54 × 10−5 2.14 × 10−5
Qe (mg/g) 92.59 73.52 69.93 90.91 71.42 94.74
R2 0.985 0.9925 0.992 0.9903 0.9902 0.997
Intra-particle diffusion model kid (g/mg/hr) 0.1146 0.1276 0.1272 0.0995 0.1222 0.1132
R2 0.996 0.992 0.948 0.9604 0.9438 0.988

and adsorbing molecules. The Temkin isotherm model states that the monolayer. The value of β parameter of Redlich-Peterson model was
heat of adsorption for the surfactant molecules in the system decreases found to be 1 and 0.817 for bentonite and sandstone samples respec-
linearly as there is an increase in surface coverage area of the adsorbent tively. These values of β also suggests that the adsorption is governed
and the adsorption process is illustrated with smooth dispersal of by Langmuir adsorption model. The maximum amount of adsorbate
binding energies of molecules, up-to maximum binding energy (Temkin adsorbed (qo) obtained from Langmuir model was found to be greatest
and Pyzhev, 1940). The adsorption data were modelled by Temkin for bentonite sample as bentonite had maximum surface area as cal-
isotherm model as represented by Eq. (6): culated from BET analysis.

qe = B ln Km + B ln Ce (6)
4.5. Gibbs free energy of adsorption
Where, B = RT/b and b represents the Temkin constant associated with
heat of sorption (J/mol), Km is the equilibrium binding constant that The Langmuir adsorption constant KL can be used to analyze the
relates to maximum binding energy (L/g), R represents the gas constant spontaneity of the adsorption by calculation of change in gibbs free
and its value is assumed to be 8.314 J/mol K, and T is the temperature energy of adsorption (ΔG). The change in gibbs free energy of adsorp-
(K). Fitting of adsorption data to Temkin model is shown in Fig. 14. tion can be calculated by Eq. (9) (Zhou and Zhou, 2014):

ΔG = -RT ln (55.5KL) (9)


4.4.4. Redlich- Peterson adsorption isotherm
The Redlich-Peterson (R-P) isotherm proposed by Redlich and Where, R is the universal gas constant (8.314 Jmol/K), T is the tem-
Peterson in 1959 is a three-parameter model that comprises of funda- perature (K), KL is the Langmuir constant (L/mol). The change in gibbs
mentals from both the Langmuir and Freundlich isotherm models (Vidal free energy was calculated to be −26.382 kJ/mol, −28.597 kJ/mol
et al., 2011; Liu et al., 2010). Redlich–Peterson equation includes 3 and −27.230 kJ/mol for sandstone, carbonate and bentonite respec-
adjustable parameters into an empirical isotherm model. The adsorp- tively. Similar results for ΔG adsorption at air-aqueous interface are
tion mechanism is distinctive and do not follow the assumption of ideal reported by other researchers (Zdziennicka and Jańczuk, 2017). The
monolayer adsorption. Experimental data obtained through adsorption negative value of ΔG indicates that the adsorption process is sponta-
study of the surfactant solutions at varying concentration was fitted neous in nature for all the samples. The values of ΔG for respective
according to Redlich-Peterson model depicted in Fig. 15. Table 2 re- samples for found to be similar indicating that each sample had similar
presents the parameters calculated from the modelling of Eq. (7): affinity for surfactant adsorption.
KCe
qe =
1 + αCβe (7) 4.6. Effect of salt content on adsorption of surfactant on adsorbent

Taking natural log on both sides, the linearized form of Eq. (7) is The presence of salt affects the physical and chemical adsorption of
presented as Eq. (8): anionic soap-nut surfactant by interfering with the electrostatic inter-
actions during the adsorption process. The effect of salt was studied at
C
ln ⎛⎜K e − 1⎞⎟ = βlnCe + lnα the surfactant concentration below and above CMC at 7000 mg/L and
⎝ qe ⎠ (8) 9000 mg/L respectively and is shown in Fig. 16. It can be seen in the
β
Where, K (L/mg) and α [(L/mg) ] are Redlich-Peterson isotherm figure that the adsorption of surfactant increases with the increase in
constants. β represents the exponential constant, with its value lying the salt concentration. Ions produced by dissolution of salt in aqueous
between 0 and 1 and helps to illustrate the adsorption isotherm model. phase are able to influence the chemisorption of ionic surfactants on
If β = 1, Eq. (7) condenses to the Langmuir model, and if β = 0, Eq. (7) solid (Chang et al., 2018). The increase in salt concentration can lead to
reduces to the linear isotherm model (Belhachem and Addoun, 2011). lead to adsorption of counterions on the surfactant micelles causing the
The adsorption pattern of soap-nut surfactant is dependent on the reduction electrostatic charge on micelle surface (Kolev et al., 2002).
interaction of surfactant molecules and the adsorbent surface. These However, this phenomenon is dominated by the suppression of elec-
interactions are in the form of various intermolecular forces existing as trical double layer of adsorbent surface. This compression of electrical
electrostatic attraction, hydrogen bonding, covalent bonding and hy- double layer of adsorbent surface leads to increase in adsorption of
drophobic bonding (Curbelo et al., 2007). For ionic surfactants, elec- surfactant with increase in salt concentration in the system (Pethkar
trostatic interaction between the surface and surfactant molecules is the and Paknikar, 1998).
major driving forces facilitating adsorption phenomenon. The charged
micelles of surfactant are attracted to the charged surface of the rock 4.7. Effect of alkali on adsorption of surfactant on adsorbent
leading to adsorption of the surfactant molecules at the surface
(Wisniewska et al., 2015). From the data presented in Table 2, it can be Alkali like sodium carbonate have been applied as sacrificial agent
seen that the R2 values of Langmuir model is closest to 1, concluding to reduce the adsorption of the surfactant in various reservoir rocks
that the adsorption of the anionic surfactant on the samples is (ShamsiJazeyi et al., 2014). Fig. 17 shows that with increase in

1280
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Fig. 20. The kinetic study for adsorption of surfactant on (a) sandstone (b) carbonate and (c) bentonite by pseudo-second order kinetics.

concentration of sodium carbonate, the adsorption of surfactant on all causes an increase of pH of the system and increases the negativity of
the samples have decreased. The decrease in adsorption of surfactant is the surface of sample. This leads to electrostatic repulsion between
due to dissociation of sodium carbonate into weak carbonic acid and surfactant molecules and surface of rock, causing a significant reduction
generation of OH− ions upon interaction with water molecules. This of surfactant adsorption. Sodium carbonate molecules also forms in-situ

1281
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

surfactants by interacting with the naphthenic acids present in the mass transfer diffusion model were calculated for surfactant con-
crude oil which facilitates in recovery of trapped oil. centrations of 7000 mg/L and 9000 mg/L and the data obtained were
analyzed and fitted with the models as depicted in Table 4. It was ob-
4.8. Effect of nanoparticle on adsorption of surfactant on adsorbent served that pseudo second order kinetic model fits well for sandstone,
carbonate and bentonite per the high values of R2 and the fitted pseudo
Figs. 18 and 19 depict the influence of alumina and silica nano- second order model have been shown for the respective samples in
particles respectively on the adsorption behavior of the soap-nut sur- Fig. 20. The value of R2 for intra-particle diffusion model also had value
factant on to the sample surfaces. Alumina nanoparticle was selected closer to 1, however, the model equation did not pass through the
for this adsorption study as it has been reported in the literature that origin point indicating that the adsorption is not governed by intra-
alumina nanoparticle is effectively able to reduce oil brine IFT and particle diffusion model.
viscosity of oil. Alumina nanoparticle aid in alteration of wettability of
intermediate oil wet rock to water wet which results good recovery of 5. Conclusions
oil. Silica nanoparticle was selected as silicon dioxide popularly called
as silica is amongst the abundant material in the earth crust and it is an Surfactant adsorption pattern was observed through experimental
essential component of sand stone and it can be easily synthesized to and modelling studies on different reservoir rock surface viz. sandstone,
nanoparticles. Researches have proved that silica nanoparticle barely carbonate and bentonite clay, considering its application for designing
show any change in the specific area when heated for temperature up to of surfactant based chemical enhanced oil recovery. Mineralogy and
650 °C proving high thermal stability (Wang et al., 1999). These morphological analysis obtained by FE-SEM, XRD and BET surface area
properties have made silica as most cost effective and commonly used measurements predicted that bentonite had more surface area available
nanoparticle. The researchers around the world have recommended for adsorption due to more porous rock surface. Zeta potential of
these nanoparticles as an EOR agent for heavy oil reservoirs. (Negin samples suspended in aqueous solution of varying pH showed that the
et al., 2016; Hendraningrat and Torsæter, 2015). surfaces were negatively charged. The presence of negatively charged
The plots clearly show that the surfactant adsorption decreases in surface and large surface area have led to the adsorption of surfactant
presence of nanoparticle. In the current study the effect of nanoparticle molecules on the sample surface. When the surfactant concentration
was studied by varying the weight percentage of nanoparticle ranging was increased in the system the adsorption onto the rock surface was
from 0.1% to 0.3% for the surfactant concentration of 7000 mg/L and also enhanced due to the more molecules available for adsorption. The
9000 mg/L. In absence of nanoparticles, the surfactant molecules are experimental data was analyzed by mathematically fitting with ad-
adsorbed on the surface of the samples. However, when nanoparticles sorption isotherms like Langmuir, Freundlich, Temkin, and Redlich-
are added to the system, it also acts as an adsorbent leading to com- Peterson models. Adsorption pattern for the surfactant-rock system was
petitive adsorption of surfactant molecules onto the surfaces of nano- best clarified by Langmuir adsorption isotherm along with the Redlich-
particles and samples (Suleimanov et al., 2011). It was observed that Peterson model that supports the Langmuir isotherm model for the
the adsorption of surfactant molecules on the nanoparticles surface was monolayer adsorption behavior. Increase in salinity was found to in-
higher due to stronger hydrogen bonding leading to less amount of crease the adsorption of surfactant molecules, however, presence of
surfactant molecules available to be adsorbed onto the rock interface. alkali and nanoparticles were found to be effective in reducing the
The bonded nanoparticles and surfactant molecules stay in the bulk of adsorption of surfactant molecules onto the rock surface, the best re-
the solution leading to reduction of surfactant adsorption on rock sur- sults were obtained using silica nanoparticles that showed reduction in
face (Yekeen et al., 2017). It is also reported that most surfactant mo- adsorption pattern. On comparing the results of adsorption study for the
lecules collide with the nanoparticles rather than interacting with rock surfactant on the different rock surfaces, it was observed that rock
surface causing reduction of surfactant adsorption (Yining et al., 2017). mineralogy plays a crucial role in surfactant adsorption. Bentonite
On comparison of data plotted in Figs. 18 and 19, it can be con- being highly porous in nature adsorbed highest amount of surfactant in
cluded that the adsorption reduction capability of silica nanoparticle is comparison of sandstone and carbonate rock surfaces. The selected
greater than the reduction capacity of alumina nanoparticle. The anionic surfactant showed least adsorption on sandstone rock surface
greater adsorption reduction capacity of silica nanoparticle can be ex- which also supports that the application of anionic surfactant in sand-
plained by almost rounded structure as compared to alumina nano- stone reservoirs is favorable for enhance oil recovery. Kinetic adsorp-
particles which have sharp edges (Yining et al., 2017). Studies by Ah- tion study for the present surfactant system was modelled according to
madi and Shadizadeh (Ahmadi and. Shadizadeh, 2013) also suggested various kinetic models and pseudo second order kinetic model was
creation of a strong hydrophobic bond between hydrophobic groups of found to be most appropriate in explaining the adsorption kinetic for
silica nanoparticles and the surfactant hydrophobic tail leading to an sandstone, carbonate and bentonite surfaces.
increased adsorption of surfactant molecules onto the silica nano-
particle surface and reduction in adsorption of surfactant molecules on Acknowledgement
the adsorbent surface.
The XRD, FE-SEM and BET surface area analysis were conducted at
4.9. Theory of adsorption kinetics of the soap-nut surfactant Central Research Facility (CRF) at Indian Institute of Technology
(Indian School of Mines), Dhanbad, Dhanbad. The present work was
Adsorption of surfactant onto the rock surface is the result of in- financed by Oil India Limited with the contract number 6206917.
teraction of solid/liquid surface and mass transfer from aqueous phase
to the rock interface. Kinetic study of adsorption helps us to predict the Appendix A. Supplementary data
adsorption rate and provide information about adsorption mechanism.
Adsorption data was analyzed by three most commonly used kinetic Supplementary data to this article can be found online at https://
models viz. pseudo-first order, pseudo second-order and intra particle doi.org/10.1016/j.petrol.2018.11.002.
diffusion model. The equations of the models and their respective
parameters are shown in Table 3. References
The above mentioned kinetic models were used to explain the ad-
sorption pattern of surfactant on to the rock surface as a function of Ahmadi, M.A., Shadizadeh, S.R., 2013. Induced effect of adding nano silica on adsorption
time. Kinetic parameters for different kinetic models which includes of a natural surfactant onto sandstone rock: experimental and theoretical study. J.
Petrol. Sci. Eng. 112, 239–247.
pseudo-first-order model, pseudo second-order model and intra-particle

1282
N. Saxena et al. Journal of Petroleum Science and Engineering 173 (2019) 1264–1283

Ahmadi, M.A., Shadizadeh, S.R., 2013. Induced effect of adding nano silica on adsorption Monfared, A.D., Ghazanfari, M.H., Jamialahmadi, M., Helalizadeh, A., 2015.
of a natural surfactant onto sandstone rock: experimental and theoretical study. J. Adsorptionof silica nanoparticles onto calcite: equilibrium, kinetic, thermodynamic
Petrol. Sci. Eng. 112, 239–247. and DLVO analysis. Chem. Eng. J. 281, 334–344.
Al-hashim, H., Obiora, V., Al-yousef, H.Y., Fernandez, F., Nofal, W., 1996. Alkaline sur- Negin, C., Ali, S., Xie, Q., 2016. Application of nanotechnology for enhancing oil
factant polymer formulation for saudi arabian carbonate reservoirs. In: Paper SPE/ recovery–A review. Petroleum 2, 324–333.
DOE 35353 Presented at the SPE/DOE Tenth Symposiumon Improved Oil Recovery, Pethkar, A.V., Paknikar, K.M., 1998. Recovery of gold from solution using
Tulsa, OK, USA. Cladosporioides biomass beads. J. Biotechnol. 63, 211–220.
Alroudhan, A., Vinogradov, J., Jackson, M.D., 2016. Zeta potential of intact natural Rahbar, M., Roosta, A., Ayatollahi, S., Ghatee, M.H., 2012. Prediction of three-dimen-
limestone: impact of potential-determining ions Ca, Mg and SO4. Colloids Surf., A sional (3-D) adhesion maps, using the stability of the thin wetting film during the
494, 83–90. wettability alteration process. Energy Fuels 26, 2182–2190.
Bai, B., Grigg, R.B., 2005. Kinetics and equilibria of calcium lignosulfonate adsorption Saha, R., Uppaluri, R.V.S., Tiwari, P., 2017. Effect of mineralogy on the adsorption
and desorption onto limestone. SPE J., SPE-93098-MS. https://doi.org/10.2118/ characteristics of surfactant-Reservoir rock system. Colloid. Surface. Physicochem.
93098-MS. Eng. Aspect. 531, 121–132.
Baik, M.H., Lee, S.Y., 2010. Colloidal stability of bentonite clay considering surface Saxena, N., Kumar, S., Mandal, A., 2018a. Adsorption characteristics and kinetics of
charge properties as a function of pH and ionic strength. J. Ind. Chem. Eng. 16, synthesized anionic surfactant and polymeric surfactant on sand surface for appli-
837–841. cation in enhanced oil recovery, Asia-Pac. J. Chem. Eng. 13 (4) e2211.
Barati, A., Najafi, A., Daryasafar, A., Nadali, P., Moslehi, H., 2016. Adsorption of a new Saxena, N., Pal, N., Ojha, K., Dey, S., Mandal, A., 2018b. Synthesis, characterization,
nonionic surfactant on carbonate minerals in enhanced oil recovery: experimental physical and thermodynamic properties of a novel anionic surfactant derived from
and modeling study. Chem. Eng. Res. Des. 105, 55–63. Sapindus laurifolius. RSC Adv. 8, 24485–24499.
Barati-Harooni, A., Najafi-Marghmaleki, A., Tatar, A., Mohammadi, A.H., 2016. ShamsiJazeyi, H., Verduzco, R., Hirasaki, G.J., 2014. Reducing adsorption of anionic
Experimental and modeling studies on adsorption of a nonionic surfactant on sand- surfactant for enhanced oil recovery: Part II. Applied aspects. Colloids Surf., A 453,
stone minerals in enhanced oil recovery process with surfactant flooding. J. Mol. Liq. 168–175.
220, 1022–1032. Sing, K.S.W., 1973. In: Everett, D.H. (Ed.), Colloid Science. Chemical Society Specialist
Belhachem, M., Addoun, F., 2011. Comparative adsorption isotherms and modeling of Periodical Report, London.
methylene blue onto activated carbons. Appl. Water Sci. 1, 111–117. Sohal, M., Thyne, A.G., Søgaard, E.G., 2016. Review of recovery mechanisms of ionically
Bera, A., Kumar, T., Ojha, K., Mandal, A., 2013. Adsorption of surfactant on sand surface modified waterflood in carbonate reservoirs. Energy Fuels 30, 1904–1914.
in enhanced oil recovery:Isotherms,kinetics and thermodynamics studies. Appl. Surf. Somasundaran, P., Hanna, H.S., 1979. Adsorption of sulfonates on reservoir rocks. SPE J.
Sci. 284, 87–99. 19, 221–232.
Bournival, G., Du, Z., Ata, S., Jameson, G., 2014. Foaming and gas dispersion properties of Standnes, D.C., Austad, T., 2000. Wettability alteration in chalk 2. Mechanism for wett-
non -ionic surfactants in the presence of an inorganic electrolyte. Chem. Eng. Sci. 116 ability alteration from oil-wet to water-wet using surfactants. J. Petrol. Sci. Eng. 28,
5 36-546. 123–143.
Chang, Z., Chen, X., Peng, Y., 2018. The adsorption behavior of surfactants on mineral Stern, Otto, 1924. Zur theorie der elektrolytischen doppelschicht. Zeitschrift für
surfaces in the presence of electrolytes– A critical review. Miner. Eng. 121, 66–76. Elektrochemie und angewandte physikalische Chemie 30 (21‐22), 508–516.
Curbelo, F.D., Santanna, V.C., Neto, E.L.B., Dutra Jr., T.V., Dantas, T.N.C., Neto, A.A.D., Suleimanov, B., Ismailov, F., Veliyev, E., 2011. Nanofluid for enhanced oil recovery. J.
Garnica, A.I., 2007. Adsorption of nonionic surfactants in sandstones. Colloids Surf., Petrol. Sci. Eng. 78, 431–437.
A 293, 1–4. Temkin, M.J., Pyzhev, V., 1940. Recent modifications to Langmuir isotherms. Acta
Douglas, H., Walker, R., 1950. The electro-kinetic behaviour of Iceland Spar against Physiochim URSS 12, 217–225.
aqueous electrolyte solutions. Trans. Faraday Soc. 46, 559–568. Vidal, C.B., Barros, A.L., Moura, C.P., A de Lima, A.C., Dias, F.S., Vasconcellos, L.C.G., A
Freundlich, H.M.F., 1906. Over the adsorption in solution. J. Phys. Chem. 57, 1100–1107. Fechine, P.B., Nascimento, R.F., 2011. Adsorption of polycyclic aromatic hydro-
Gogoi, S.B., 2009. Adsorption of non-petroleum base surfactant on reservoir rock. Curr. carbons from aqueous solutions by Modified Periodic Mesoporous Organosilica. J.
Sci. 97, 1059–1063. Colloid Interface Sci. 357, 466–473.
Gregersen, C.S., Mahdi, K., Vladimir, A., 2013. ASP design for the Minnelusa formation Wang, L., Wang, Z., Yang, H., Yang, G., 1999. The study of thermal stability of the SiO2
under low-salinity conditions: impacts of anhydrite on ASP performance. Fuel 105, powders with high specific surface area. Mater. Chem. Phys. 57, 260–263.
368–382. Wang, H., Zhou, A., Peng, F., Yu, H., Yang, J., 2007. Mechanism study on adsorption of
Gupta, R., Mohanty, K., 2007. Temperature effects on surfactant-aided imbibition into acidified multi-walled carbon nanotubes to Pb (II). J. Colloid Interface Sci. 316
fractured carbonates. SPE J. 15, 588–597. 277–28.
Helmholtz, H., 1853. Ueber einige Gesetze der Vertheilung elektrischer Ströme in Wang, J., Han, M., Fuseni, A., Fuseni, A.B., Cao, D., 2013. Surfactant adsorption in sur-
körperlichen Leitern mit Anwendung auf die thierisch-elektrischen Versuche. factant-polymer flooding for carbonate reservoirs. In: SPE Middle East Oil and Gas
Annalen der Physik und Chemie 165, 211–233 (in German). Show and Conference.
Hendraningrat, L., Torsæter, O., 2015. Metal oxide-based nanoparticles: revealing their Wang, J., Ming, H., Alhasan, B.F., Dongqing, C., 2015. Surfactant adsorption in surfac-
potential to enhance oil recovery in different wettability systems. Appl. Nanosci. 5, tant-polymer flooding for carbonate reservoirs. In: SPE Middle East Oil & Gas Show
181–199. and Conference. Society of Petroleum Engineers.
Hiorth, A., Cathles, L., Madland, M., 2010. The impact of pore water chemistry on car- Weber, W.J., Morriss, J.C., 1963. Kinetics of adsorption on carbon from solution. J.
bonate surface charge and oil wettability. Transport Porous Media 85, 1–21. Sanitary Eng. Div. Am. Soc. Civ. Eng. 89, 31–60.
Hirasaki, G., 1991. Fundamentals and surface forces. SPE Form. Eval. 6, 217–226. Wisniewska, M., Chibowski, S., Urban, T., 2015. Adsorption properties of the nano-
Hirasaki, G.J., Miller, C.A., Puerto, M., 2008. Recent advances in surfactant EOR. In: SPE zirconia/anionic polyacrylamide system minuseffects of surfactant presence, solution
Annual Technical Conference and Exhibition, Denver, USA, vol. 2008. pH and polymer carboxyl groups content. Appl. Surf. Sci. 370, 351–356.
Hou, B.F., Wang, Y.F., Huang, Y., 2015. Mechanistic study of wettability alteration of oil- Yan, L.M., Li, Y.L., Cui, Z.G., Song, B.L., Pei, X.M., Jiang, J.Z., 2017. Performances of
wet sandstone surface using different surfactants. Appl. Surf. Sci. 330, 56–64. guerbet alcohol ethoxylates for Surfactant−Polymer flooding free of alkali. Energy
Huang, Y.C., Fowkes, F.M., Lloyd, T.B., Sanders, N.D., 1991. Adsorption of calcium ions Fuels 31, 9319–9327.
from calcium chloride solutions onto calcium carbonate particles. Langmuir 7, Yekeen, N., Idris, A.K., Manan, M.A., Samin, A.M., Risal, A.R., Kun, T.X., 2016. Bulk and
1742–1748. bubble scale experimental studies of influence of nano-particle on foam stability.
Kolev, V.L., Danov, K.D., Kralchevsky, P.A., Broze, G., Mehreteab, A., 2002. Comparison Chin. J. Chem. Eng. 25, 347–357.
of the van der Waals and frumkin adsorption isotherms for sodium dodecyl sulfate at Yekeen, N., Manan, M.A., Idris, A.K., Samin, A.M., Risal, A.R., 2017. Experimental in-
various salt concentrations. Langmuir 18, 9106–9109. vestigation of minimization in surfactant adsorption and improvement in surfactant-
Lagergren, S., 1898. Zur theorie der sogenannten adsorption gelöster stoffe. K. - Sven. foam stability in presence of silicon dioxide and aluminum oxide nanoparticles. J.
Vetenskapsakademiens Handl. 24, 1–39. Petrol. Sci. Eng. 159, 115–134.
Langmuir, I., 1916. The constitution and fundamental properties of solids and liquids. Yining, W., Chen, W., Dai, C., Huang, Y., Li, H., Zhao, M., He, L., Jiao, B., 2017. Reducing
Part I. Solids. J. Am. Chem. Soc 38, 2221–2295. surfactant adsorption on rock by silica nanoparticles for enhanced oil recovery. J.
Liu, Q.S., Zheng, T., Wang, P., Jiang, J.P., Li, N., 2010. Adsorption isotherm, kinetic and Petrol. Sci. Eng. 153, 283–287.
mechanism studies of some substituted phenols on activated carbon fibers. Chem. Zdziennicka, A., Jańczuk, B., 2017. Thermodynamic parameters of some biosurfactants
Eng. J. 157, 348–356. and surfactants adsorption at water-air interface. J. Mol. Liq. 243, 236–244.
Michalske, T.A., Freiman, S.W., 1983. A molecular interpretation of stress corrosion in Zhou, X., Zhou, X., 2014. The unit problem in the thermodynamic calculation of ad-
vitreous silica. J. Am. Ceram. Soc. 66, 284–288. sorption using the Langmuir equation. Chem. Eng. Commun. 201, 1459–1467.

1283

You might also like