Anda di halaman 1dari 13

Applied Surface Science 284 (2013) 87–99

Contents lists available at ScienceDirect

Applied Surface Science


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

Adsorption of surfactants on sand surface in enhanced oil recovery:


Isotherms, kinetics and thermodynamic studies
Achinta Bera, T. Kumar, Keka Ojha, Ajay Mandal ∗
Department of Petroleum Engineering, Indian School of Mines, Dhanbad 826004, India

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

Article history: Adsorption of surfactants onto reservoir rock surface may result in the loss and reduction of their concen-
Received 15 May 2013 trations in surfactant flooding, which may render them less efficient or ineffective in practical applications
Received in revised form 5 July 2013 of enhanced oil recovery (EOR) techniques. Surfactant flooding for EOR received attraction due to its abil-
Accepted 7 July 2013
ity to increase the displacement efficiency by lowering the interfacial tension between oil and water and
Available online 16 July 2013
mobilizing the residual oil. This article highlights the adsorption of surfactants onto sand surface with
variation of different influencing factors. It has been experimentally found that adsorption of cationic
Keywords:
surfactant on sand surface is more and less for anionic surfactant, while non-ionic surfactant shows inter-
Surfactant
Adsorption isotherm
mediate behaviour. X-ray diffraction (XRD) study of clean sand particles has been made to determine the
Isotherm model main component present in the sand particles. The interaction between sand particles and surfactant has
Adsorption kinetics been studied by Fourier Transform Infrared (FTIR) Spectroscopy of the sand particles before and after
Thermodynamics of adsorption aging with surfactant. Salinity plays an important role in adsorption of anionic surfactant. Batch experi-
ments were also performed to understand the effects of pH and adsorbent dose on the sorption efficiency.
The sand particles exhibited high adsorption efficiency at low pH for anionic and nonionic surfactants.
But opposite trend was found for cationic surfactant. Adsorption data were analyzed by fitting with Lang-
muir, Freundlich, Redlich-Peterson, and Sips isotherm models. Results show that the Langmuir isotherm
and pseudo-second order kinetics models suit the equilibrium and kinetics of adsorption on sand surface.
Thermodynamics feasibility of the adsorption process was also studied to verify the spontaneity of the
process.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction decades. A number of studies have been conducted on the adsorp-


tion of ionic and nonionic surfactants onto reservoir rocks [8–22].
Adsorption of surfactants on rock/clay/sediment solid matrix The solid surfaces are either positively or negatively charged in
may result in the loss and reduction of their concentrations, which the aqueous medium by ionization/dissociation of surface groups or
may render them less efficient or ineffective in practical appli- by the adsorption of ions from solution onto a previously uncharged
cations of EOR techniques. Surfactants are also widely used in surface. At low surfactant concentrations, the charge on the elec-
various industrial processes for their favourable physicochemical trical double layer (proposed by Helmholtz in 1879, and modified
characteristics like detergency, foaming, emulsification, dispersion by Stern in 1924) of the solid surface largely determines the sur-
and solubilization effects [1–4]. Due to extreme ability to reduce factant adsorption. The surfactant molecules are adsorbed on rock
oil-water interfacial tension (IFT), surfactants are very important surface or sediments as a single monomer and form monomeric
materials in chemical flooding for EOR methods. Adsorption of layer at low concentration of surfactant solution. As the surfac-
surfactants from aqueous solutions in porous media is a funda- tant concentration increases, the adsorbed surfactant monomers
mental issue in EOR from oil reservoirs because surfactant loss due tend to aggregate and form micelles [13,23]. This aggregate can
to adsorption on the reservoir rocks impairs the effectiveness of form one layer (ad micelles) or two layer (hemi micelles). The
the chemical solution injected to reduce the IFT of oil-water and onset of hydrophobic interaction between the adsorbed surfactant
may turn into the process economically unfeasible [5–8]. Surfac- molecules leads to a substantial increase in the adsorption level-
tant adsorption at solid/liquid interface has been studied for several ling off at the critical micelle concentration (CMC) [24,25]. In order
to lower the adsorption, negatively charged surfactants are usually
considered as the main surfactant species of the slug and so anionic
surfactants are believed to be the most used type of chemicals in the
∗ Corresponding author. Tel.: +91 326 2235485; fax: +91 326 2296632. flooding of sandstone oil reservoirs [26]. The adsorption of surfac-
E-mail address: mandal ajay@hotmail.com (A. Mandal). tants from the solution is affected by its physicochemical properties

0169-4332/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apsusc.2013.07.029
88 A. Bera et al. / Applied Surface Science 284 (2013) 87–99

interface at rock-fluid boundary. This process can be explained as


Nomenclature the interface is energetically favoured by the surfactant molecules
compared to the bulk phase [35,36]. It has been shown that the
EOR enhanced oil recovery
nature of the adsorption isotherm depends to a large extent on
XRD X-ray diffraction
the type of surfactant used, the morphological and mineralogical
FTIR Fourier Transform Infrared
characteristics of the rock, and the type of electrolytes present in
IFT interfacial tension
solution [37]. The adsorption of surfactants can be affected by the
CMC critical micelle concentration
surface charge on the rock surface and fluid interfaces [38–41]. Pos-
SDS sodium dodecylsulphate
itively charged cationic surfactant is attracted to negatively charged
CTAB cetyltrimethylammonium bromide
surfaces, while negatively charged anionic surfactant is attracted to
DOE design of experiments
positively charged surfaces. The salinity and pH of brine strongly
COD chemical oxygen demand
affect the surface charge. When the effects of brine chemistry are
MSE Mean Square Error
removed, silica tends to adsorb simple organic bases (cationic sur-
MINAPE Minimum absolute percentage error
factant), while the carbonates tend to adsorb simple organic acids
MAXAPE Maximum absolute percentage error
(anionic surfactant). This occurs because silica normally has a neg-
atively charged weak acidic surface in water near neutral pH, while
Variables
the carbonates have positively charged weak basic surfaces. Loss
C0 Initial concentrations of surfactants (mg/g)
of surfactants owing to their interactions with reservoir rocks and
Ce Equilibrium aqueous concentration of surfactants
fluid is possibly the most important factor that can determine the
(mg/L)
efficiency of a micellar flooding process [42].
V Volume of the surfactant solution (L)
Studies of adsorption kinetics and equilibrium of different
m Weight of the sand particles (g)
surfactants are very practical tests in laboratory for study of sur-
RL Separation factor or equilibrium parameter
factant adsorption onto rock surface. These phenomena depend
KR Redlich-Peterson isotherm constants (L/mg)
on the nature of the surfactants and also the solid-liquid interface
KS Sips isotherm constant [(L/mg)m s ]
[36,43–45]. Recently Ahmadi et al. [46] have studied the adsorp-
ˇ Exponential factor
tion behavior of the Glycrihiza Glabra, a novel nonionic surfactant,
KL Langmuir equilibrium constant (L/mg)
onto carbonate rock and Ahmadi and Shadizadeh [47] have inves-
KF Freundlich adsorption constants related to sorption
tigated the effect of nanosilica on adsorption behavior of Zyziphus
capacity (mg/g)
Spina Christi onto rock surface. Ahmadi et al. [46] concluded that
t Amount of surfactants adsorbed on sand particles at
adsorption isotherm follows the Langmuir model. On the other
time t (mg/g)
hand when nanosilica is used the Linear, Langmuir, and Temkin
k1 Rate constant of the pseudo-first-order adsorption
equilibrium adsorption models were not suitable for predicting the
(min−1 )
surfactant adsorption, but the Freundlich equilibrium adsorption
k2 Rate constant of the second-order equation
was in good agreement between the experimental data. They also
(g/mg/min)
studied the kinetics of the adsorption and showed that the pro-
G◦ Change in Gibbs energy (J/mol)
cess follows the second order kinetic model. Gogoi [48] reported
S◦ Change in entropy (J/mol)
the effect of NaCl concentration and pH on the adsorption equilib-
H◦ Change in enthalpy (J/mol)
rium of Na-lignosulfonate onto reservoir rocks. He demonstrated
 Amount of adsorbate adsorbed (mg/g)
that adsorption increases with increasing NaCl concentration but
 max Maximum amount adsorbed (mg/g)
decreases with increasing pH.
n Sorption intensity
The net adsorption of surfactant in an EOR process strongly
˛R Redlich-Peterson isotherm constants [(L/mg)ˇ ]
depends on the presence of oil and the flow field. When a surfac-
ms Empirical constant in Sips isotherm
tant slug is injected as displacing fluid, it undergoes partitioning
T Temperature (K)
into oil and water and lowers the interfacial tension between oil
R2 Regression coefficient
and water thereby increasing the capillary number. As a result, the
R Universal Gas Constant (8.314 J/K/mol)
trapped immobile oil becomes mobile. At the same time, an oil-
Kid Rate constant of intraparticular diffusion
in-water emulsion is formed which blocks the larger pores leading
(mg/g/min)
to an improvement in the effective mobility ratio. Otherwise the
C Intercept
injected surfactant solution flows through the highly permeable
zone bypassing the trapped oil in smaller pores. The injected sur-
factant continues to mobilize oil, until the surfactant is diluted or
such as pH [27–29], temperature [30,31], ionic strength [27,31], otherwise lost due to adsorption on the rock surface. Consequently
adsorbent dose [32] and electrolyte concentration [27,30,33]. These the surfactant solutions with lower concentration could not be able
physicochemical properties of solutions can also influence in the to lower the interfacial tension and mobilize oil. At that point, the
dissolution behavior of minerals resulting significant changes in process degenerates into a water flood. Hence to design a surfactant
the precipitation behavior of the surfactants [34]. A slight variation flooding for EOR, it is very important to have a complete knowledge
in one of the above factors or the other can result in a significant of adsorption of the specific surfactant on the reservoir rock under
change in the adsorption characteristics of the system. the reservoir conditions.
Adsorption is a unit operation in which dissolved constituents In the present paper the adsorptions of three different surfac-
are removed from the solvent by interphase transfer to the sur- tants namely anionic, cationic, and nonionic by clean sand particles
face of an adsorbent particle. In chemical flooding, surfactants have been investigated with variation of different parameters i.e.,
are inevitably adsorbed on the surface of reservoir rock by the salinity, pH, temperature, and adsorbent dose. Adsorption data
rock/oil/brine interaction. Surfactant adsorption in porous media is have been analysed by fitting with Langmuir, Freundlich, Redlich-
a typically complex phenomenon (e.g., mass transfer and reaction). Peterson, and Sips isotherm models. Kinetics of adsorption has also
Adsorption in porous media is a phenomenon in which trans- been carried out with anionic surfactant. Thermodynamic feasibil-
port of surfactant molecules takes place from bulk phase onto the ity of the adsorption process has also been studied to verify the
A. Bera et al. / Applied Surface Science 284 (2013) 87–99 89

Fig. 1. Molecular structures of the surfactants used in the present study.

spontaneity of the process. Semi-quantitative analysis of clean sand water for several times followed by settling and decanting. After
particles has been done by X-ray diffraction (XRD) study. removing the dust particles the residual wet sand particles were
dried at 353 K for 18 h. The clean dried sand particles were used for
2. Experimental the experimental purposes.

2.1. Materials used 2.3.2. XRD study of clean sand powder


The clean sands were ground to prepare powder sample. X-ray
Three different categories of surfactants such as anionic, cationic diffractogram of prepared sample were recorded in a wide range
and nonionic were used to determine the adsorption isotherms of Bragg angle 2 (10◦ ≤ 2 ≤ 90◦ ) using Bruker D8 advanced XRD
on the clean sand particles (60–70 mesh size). Anionic surfactant, measuring instrument with Cu target radiation ( = 0.154056 nm).
Sodium dodecylsulphate (SDS) (with 98% purity) was purchased The data were analysed with the help of the JCPDS files.
from Fisher Scientific, India and cationic surfactant, Cetyltrimethy-
lammonium bromide (CTAB) of 98% pure was procured from Merck, 2.3.3. FTIR study
India, both were used in the present study. Tergitol 15-S-7 (99.5% The apparatus used for measuring the FTIR spectra of the sand
pure) from sigma-Aldrich, Germany was used as nonionic surfac- particles before and after surfactant treatment in the range of
tant. The molecular structures of the surfactant have been given in 450–4000 cm−1 , was a PerkinElmer Spectrum version 10.03.07 FTIR
Fig. 1. Sodium Chloride (NaCl) procured from Qualigens Fine Chem- spectrometer. The instrument is operated by Spectrum two soft-
icals, India, was used for preparation of brine. Reverse osmosis ware supplied by PerkinElmer (USA). For the FTIR analysis, 4 mg of
water from Millipore water system (Millipore SA, 67120 Molshein, dried sample was mixed with potassium bromide (KBr) (∼300 mg),
France) was used for preparation of solutions. which was used as a reference standard sample. The mixture was
compressed by hydraulic pump to prepare pallet and the pallet was
2.2. Design of experiments (DOE) placed in a desiccator to remove moisture content of the sample.
The dried sample then was used for experimental purpose.
DOE refers to the process of planning, designing and analyz-
ing the experiment so that valid and objective conclusions can 2.3.4. Determination of critical micelle concentration (CMC)
be drawn effectively and efficiently. Three different surfactants Measurement of surface tension is very much useful supple-
namely SDS (anionic), CTAB (cationic) and Tergitol 15-S-7 (non- mentary test method for determination of CMC of surfactant. It is
ionic) have been used for the adsorption studies at different salinity, particularly useful when only very small quantities of an experi-
pH, temperature and adsorbent dose. Semi-quantitative analysis mental surfactant are available. In the present study surface tension
of clean sand particles has been done by X-ray diffraction (XRD) of the different concentrated surfactant solutions were measured
study to determine the main component present in the sand by a programmable tensiometer (Kruss GmbH, Germany, model:
particles. CMCs of the surfactants were determined by surface ten- K20 EasyDyne) under atmospheric pressure by the Du Noüy ring
sion method. Adsorption data have been analysed by fitting with method. CMCs of the surfactants were determined from plot of
Langmuir, Freundlich, Redlich-Peterson, and Sips isotherm models. surface tension and surfactant concentration. The concentration at
Kinetics of adsorption has also been carried out with anionic sur- the inflexion point of the curve is termed as CMC. During the mea-
factant. Effect of temperature on surfactant adsorption has been surement, the experimental temperature was maintained at 298 K.
investigated. Thermodynamic feasibility of the adsorption process The platinum ring was thoroughly cleaned with acetone and flame-
has also been investigated to verify the spontaneity of the process. dried before each measurement. In all cases the standard deviation
did not exceed ±0.1 mN/m.
2.3. Experimental procedures
2.3.5. Adsorption isotherms
2.3.1. Preparation of clean sand particles (adsorbent) A series of batch experiments were carried out to determine
Sands which are used for making building were first sieved to get the adsorption isotherms of different types of surfactants on the
60–70 mesh sized sand particles and washed with double distilled adsorbent. 8 g of clean sand particles were added to a set of 50 ml
90 A. Bera et al. / Applied Surface Science 284 (2013) 87–99

surfactant solutions in a 100 ml glass vials and allowed to con-


duct the experiments by constant shaking at 303 K for 24 h on
a temperature controller horizontal shaker machine (Model No.
NovaShake BB03) at 120 rpm speed. After adsorption, the surfac-
tant solutions were isolated by centrifugation with Remi centrifuge
instrument (Model No. Remi R-8 C). The equilibrium concentration
of the surfactant solutions were determined by Chemical Oxygen
Demand (COD) measurement of the solution. The amount of sur-
factant adsorbed on the adsorbent,  (mg/g), was calculated by a
mass balance relation (1):
V
 = (C0 − Ce ) (1)
m
where, C0 and Ce are the initial and equilibrium concentrations
of surfactants (mg/g) respectively, V is the volume of the surfactant
solution (L), and m is the weight of the sand particles (g) (adsorbent)
used.
The effects of the pH, temperature, NaCl concentrations and
adsorbent dose on the adsorption capacity of the adsorbent to
the anionic surfactant, SDS were also investigated. To adjust the
required pH values of the solutions, HCl (0.1 N) and NaOH (0.1 N) Fig. 2. XRD study of the crushed sand particles used in the present work.
solutions were used. The thermodynamic study has been conducted
by change in temperature with above mentioned speed and proce- 2851.85 cm−1 and 2924.64 cm−1 . These two peaks indicates the
dure. symmetric and asymmetric –CH2 stretching.
In Fig. 3(b) the results of the sample treated with SDS has
2.3.6. Adsorption kinetics and thermodynamics been shown. In general, SDS exhibits bands due to symmetric
8 g of clean sand particles were put into 50 ml of SDS solutions at and asymmetric stretching and deformation of methylene chain at
three different concentrations of 400 ppm, 800 ppm, and 1000 ppm 2851.85 cm−1 and 2922 cm−1 . The 2851.85 cm−1 and 2920.73 cm−1
respectively. The adsorption kinetics experiments were carried out bands overlapped with pure sand peaks. It is also seen that the sam-
at 303 K and the concentration of SDS in the solutions were deter- ple treated with SDS, the stretching vibration of the S O bond is
mined at regular intervals until an equilibrium concentration was observed at 1360.23 which is overlapped with different small peaks
achieved. of pure sand. The stretching vibration of alkyl C H bond is indicated
The effect of temperature (thermodynamic study) was carried in SDS treated sand as a strong and sharp peak at 2920.73 cm−1 ,
out by shaking 8 g of clean sand particles in 50 ml surfactant solu- which shows that SDS is adsorbed on the sand surface.
tion at different temperatures (303, 313, and 323 K) in temperature Fig. 3(c) shows the results of the sample treated with CTAB. The
controlled shaker. After 6 h, the sample was centrifuges and the CH2 group to peak at 1637.18 cm−1 for SDS treated sand shifted
concentrations of the solutions were determined. to 1627.98 cm−1 and therefore CTAB adsorption on sand particles
also takes place. The intense bands near 2850.52 and 2919.33 cm−1
3. Results and discussion can be assigned to the C H stretching and deformation vibrations
of CTAB which are overlapped with spectra of pure sand. The shif-
3.1. Characterization of used sand particles and their interaction ting of methylene chain at 2850.52 cm−1 and 2919.33 cm−1 from
with surfactant 2852.1 cm−1 is due to adsorption of CTAB on sand surface in solu-
tion phase.
3.1.1. XRD study In Fig. 3(d) FTIR spectra of Tergitol 15-S-7 treated sand is pre-
The sand particles have been characterized by XRD study. Fig. 2 sented. In this case C H stretching vibration of surfactant shows at
shows the X-ray diffractogram of the powder sample. The single 2927.16 cm−1 instead of 2920 cm−1 . This is due to in solution Ter-
headed peak indicates that there is no impurity in the sample and gitol 15-S-7 with ethoxylated group gets adsorbed on sand surface
only one phase is present. The characteristic peaks are obtained and absorption band is shifted. An additional band at 1888.05 cm−1
at 21◦ , 27.74◦ , 28.57◦ , 47.23◦ , 60◦ and 76◦ , etc. The main peak is appeared due to adsorption of this ethoxylated nonionic surfac-
was obtained at 27.47◦ . JCPDS (file no. 861630) record indicates tant.
the presence of silica in the pure sand. The other peaks show the In all cases after treatment with surfactants, bands at
presence of quartz in low quantity. 3468.68 cm−1 and 1637.18 cm−1 , 3468.86 cm−1 and 1627.98 cm−1 ,
and 3469.04 cm−1 and 1626.73 cm−1 for SDS, CTAB, Tergitol 15-S-7
3.1.2. FTIR study of sand particles respectively originated from the stretching vibration of OH group
The main application of this technique is to detect the struc- of interlayer water molecule during surfactant adsorption.
ture of chemical species and provide qualitative measurement,
based on the adsorption and molecular vibration peaks. The results 3.2. Critical micelle concentration and effectiveness of the
of the FTIR test for pure sand before and after treatment with surfactants
different surfactants are presented in Figs. 3(a)–(d). The infrared
spectra of pure sand shows adsorption peaks at 776.33 cm−1 It is well known that the surfactants reduce the surface ten-
and 1080.17 cm−1 , in the region of stretching vibration for Si- sion of water by getting adsorbed on the liquid–gas interface. The
O symmetric and asymmetric bond vibration respectively. Again critical micelle concentration (CMC), one of the main parameters
absorption bands at 521.27 cm−1 , 693.91 cm−1 are related to the for surfactants, is the concentration at which surfactant solutions
bending vibration of Si-O group in asymmetric and symmetric begin to form micelles in large amount [49]. Surface tensions of
vibration. So it is clear that the used sand particles contain pure the above three surfactants (SDS, CTAB, and Tergitol 15-S-7) solu-
silica as main composition. The sand sample also shows peaks at tions at different concentrations were measured and plotted as a
A. Bera et al. / Applied Surface Science 284 (2013) 87–99 91

Fig. 3. FTIR spectra of sand particles before and after surfactant treatment in brine (2 wt% NaCl): (a) pure sand; (b) treated with SDS; (c) treated with CTAB; (d) treated with
Tergitol 15-S-7.

function of concentration in Fig. 4. The concentration at the inflex- parameter, and it is represented by the following equation [51,52]:
ion point of the curve is critical micelle concentration. The lowest
surface tension value achieved by Tergitol 15-S-7 is 30 mN m−1 1
which is significantly lower than the surface tension value of water. RL = (3)
1 + KL C0
The CMCs of the surfactants are found to be 0.23 wt%, 0.0345 wt%,
and 0.0051 wt% for SDS, CTAB, and Tergitol 15-S-7 respectively. where, C0 (mg/L) expresses initial adsorbate concentration in aque-
ous solution. KL (L/mg) is the Langmuir constant. The RL parameter
gives important signs on the compatibility of adsorption for the
selected adsorbent–adsorbate pair. There are four possibilities for
3.3. Adsorption isotherms of surfactants on sand particles
the RL value:
The Langmuir adsorption isotherm and the Freundlich adsorp-
• In the case 0 < RL < 1, adsorption is favorable.
tion isotherm are two common isotherms used to describe the
• In the case RL > 1, adsorption is unfavorable.
equilibrium adsorption isotherm. Another two isotherms such as
• RL = 1 indicates linearity of adsorption.
Redlich-Peterson and Sips are considered here to describe the
• In the case RL = 0, adsorption is irreversible.
experimental data and find out the best fitted model for adsorption
of surfactant on sand surface.
The Langmuir equation relates the amount of solid adsorbate The values of RL obtained in this study were between 0.0445
adsorbed,  , to the equilibrium liquid concentration at a fixed tem- to 0.3507, indicating that the adsorption of surfactant onto sand
perature. The equation was developed by Irving Langmuir [50] in surface is favourable.
1916 and is expressed in this nonlinear form as follows: The Freundlich isotherm assumes that if the concentration of the
solute in the solution at equilibrium, Ce , is raised to the power 1/n,
1/n
C
max KL Ce the amount of the solute adsorbed being  , the e is constant
 = (2) at given temperature and the nonlinear form of the equation is
1 + KL Ce
expressed as:
where,  is the amount of adsorbate adsorbed (mg/g);  max is the 1/n
 = KF Ce (4)
maximum amount adsorbed (mg/g); KL is the Langmuir equilib-
rium constant (L/mg); Ce is the equilibrium aqueous concentration where, KF (mg/g) and n are the Freundlich adsorption constants
(mg/L). It is well-known that the Langmuir isotherm is applicable related to sorption capacity and sorption intensity, respectively.
for monolayer adsorption because of the homogeneous surface of Freundlich isotherm has been derived by assuming an exponen-
a finite number of identical sites. Another important parameter of tially decaying sorption site energy distribution. The Freundlich
the Langmuir isotherm model is the term “RL ” which is a nondi- isotherm assumes that the surfactant adsorption occurs on a het-
mensional constant and called as separation factor or equilibrium erogeneous surface by multilayer sorption.
92 A. Bera et al. / Applied Surface Science 284 (2013) 87–99

Fig. 4. Plot of surface tension vs. surfactant concentration for finding the CMCs of the surfactants: (a) SDS, (b) CTAB and (c) Tergitol 15-S-12.

The Freundlich constant (1/n) is related to the adsorption inten- where,  max (mg/g) is the maximum monolayer surfactant adsorp-
sity of the adsorbent. When, 0.1 < 1/n < 0.5, adsorption is favourable; tion capacity on sandstone. KS [(L/mg)m s ] is the Sips isotherm
0.5 < 1/n ≤ 1, it is easy to adsorb; 1/n > 1, it is difficulty to adsorb [53]. constant representing the energy of adsorption. ms is the empir-
The Redlich-Peterson isotherm model is used as a compromise ical constant. The values of KS ,  max and m can be obtained by
between Langmuir and Freundlich isotherm models [54,55]. The
non-linearized form of Redlich-Peterson isotherm model can be
Table 1
given as follows:
Adsorption isotherm parameters of the surfactants.
KR Ce
 = ␤
(5) Isotherm models Parameters Surfactants
1 + ˛RCe SDS CTAB Tergitol 15-S-7

where, KR (L/mg) and ˛R [(L/mg)ˇ ]


are Redlich-Peterson isotherm Langmuir  max (mg/g) 0.771 0.867 0.816
constants. ˇ is the exponent which lies between 0 and 1 and can KL × 102 (L/mg) 2.122 1.851 2.152
R2 0.975 0.968 0.958
characterize the adsorption isotherm. If ˇ = 1, Eq. (5) reduces to the
Langmuir isotherm model, and if ˇ = 0, Eq. (5) reduces to the linear Freundlich KF (mg/g) 0.225 0.302 0.256
isotherm model. The values of KR , ˛R and ˇ can be obtained by 1/n 0.177 0.149 0.171
R2 0.957 0.812 0.908
nonlinear regression method and are listed in Table 1.
The Sips isotherm model is obtained by introducing a power law Redlich-Peterson KR × 102 (L/mg) 2.626 4.524 3.162
˛R × 102 (L/mg)ˇ 6.244 8.826 7.246
expression of the Freundlich isotherm into the Langmuir isotherm
ˇ 0.914 0.851 0.827
[56]. The non-linearized form of Sips isotherm model can be given R2 0.936 0.797 0.889
as follows:
Sips  max (mg/g) 0.901 1.877 1.737
max KS Cems KS × 102 (L/mg)m s 9.302 15.991 6.752
 = (6)
1 + KS Cems ms 0.568 0.237 0.209
R2 0.943 0.799 0.899
A. Bera et al. / Applied Surface Science 284 (2013) 87–99 93

The equilibrium amount of surfactant adsorbed on the sand


particles depends on their structures and nature of head groups.
It is clear from Fig. 5 that the amount of SDS adsorbed on the
adsorbent shows the lowest value compared to the others. The
equilibrium amount of CTAB and Tergitol 15-S-7 adsorbed on sand
particles are considerably higher than SDS. For all the surfac-
tants, it was found that there is a sudden increase in adsorption
isotherm as concentration of the surfactant increases. The sud-
den increase in adsorption isotherm may be described in terms
of formation of surface aggregates, known as “hemi micelles” of
the surfactant molecules on the sand surface due to lateral inter-
action between hydrocarbon chains. This lateral attraction force
generates an additional driving force, which superimposes existing
electrostatic attraction causing a sharp increase in adsorption. In all
cases the increase of adsorption with concentration up to a certain
point and then no increase have been observed. In case of CTAB
when surfactant concentration reaches CMC, micelles starts to form
and exist in the bulk solution and act as chemical potential sink
for additional surfactant added to the system. As a result, surfac-
Fig. 5. Adsorption isotherms of different types of surfactants at 303 K. tants cannot adsorb onto the surface and plateau of the adsorption
isotherm shown in Fig. 5 is characterized by little or no increases
in surfactant adsorption with increasing surfactant concentration.
nonlinear regression method and are given in Table 1. If the value With increase in SDS concentration strong repulsion takes place
of KS approaches 0, the Sips isotherm will become a Freundlich between sand surface and surfactant molecules due to negative
isotherm. While the value of ms = 1 or closer to 1, the Sips isotherm head groups of SDS surfactant. Therefore before CMC no increase
equation reduces to the Langmuir equation; that is, adsorption in adsorption also takes place with increasing concentration of
takes place on homogeneous surface [57]. surfactant. In case of Tergitol 15-S-7 after CMC small increase of
Fig. 5 shows the adsorption of different types of surfactants adsorption takes place due to weak hydrophobic and H-bond inter-
on sand surfaces at 303 K. To quantify the adsorption capacity of action.
the sand particles for surfactant adsorption, Langmuir Freundlich, The adsorption of an ionic surfactant at solid–liquid interface is
Redlich-Peterson, and Sips adsorption isotherm models have been strongly influenced by the compositions of the sand which makes
used. Fig. 6 depicts the different adsorption models for SDS sur- the sand surface negatively charged therefore, weak interaction
factant. Curve fittings for Langmuir, Freundlich, Redlich-Peterson, takes place with anionic surfactant (SDS) having their negatively
and Sips adsorption isotherm models for the other two surfactants charged head part. So the SDS adsorption capacity on sand parti-
like CTAB and Tergitol 15-S-7 have been given in Figs. S1 and S2 cles is not significantly high. However, CTAB is a cationic surfactant,
(supplementary information) respectively. The calculated results and the adsorption takes place mainly due to presence of some
from the curves of Langmuir, Freundlich, Redlich-Peterson, and Sips charged components of sand particles such as silica which are neg-
isotherm adsorption models for the surfactants have been sum- ative in nature at neutral pH or in water. The high adsorption
marized in Table 1. The values of regression coefficient (R2 ), mean capacity of CTAB on sand particles may be explained on the basis
square error (MSE), minimum absolute percentage error (MINAPE), of electrostatic interaction that exists between negatively charged
and maximum absolute percentage error (MAXAPE) indicate that adsorbent and positively charged head group of surfactant. Adsorp-
the Langmuir model is well fitted with the adsorption isotherm of tion of nonionic surfactant occurred on solid adsorbent due to weak
the surfactants on sand surface. The details of these values have hydrophobic and hydrogen bond interactions between surfactants
been given in supplementary information (Table S1). and the adsorbent. Since no positive and negative charge can exist
on nonionic surfactants so the adsorption capacity of Tergitol 15-
S-7 is also low.

3.4. Effect of salt concentration on adsorption isotherm of SDS

Adsorption isotherms for SDS surfactant solution at different


salinities have been shown in Fig. 7. At the interface between sur-
factant and sand particles, there is always an unequal distribution
of electrical charges. This unequal charge distribution gives rise
to a potential across the interface and forms a so-called electrical
double layer [58]. With increase in NaCl concentration, the elec-
trical double layer on the surface of adsorbent is compressed and
electrostatic repulsion between the adsorbed surfactant species
decreases, which results in the increase of adsorption capacity.
The surfactant adsorption capacity increases with the increase
in salinity of the system at a constant temperature of 303 K.
These facts imply that the adsorption of SDS on sand particle
adsorbent is favored at high salinity and therefore the adsorption
process is seen to be a chemical process with increasing salin-
ity.
Fig. 6. Different isotherm models fit for adsorption of SDS surfactant on sand parti-
The curve fittings for Langmuir and Freundlich adsorption
cles at 303 K. isotherm models have been depicted in Figs. 8a and b respectively
94 A. Bera et al. / Applied Surface Science 284 (2013) 87–99

Fig. 7. Adsorption isotherm of SDS surfactant on sand surface at different salinities


Fig. 9. The effect of the amount of sand on the adsorption process of the surfactants.
of brine at 303 K.

3.5. Effect of adsorbent dose on the extent of surfactants


at 303 K. Table 2 shows the parameters obtained the two mod- adsorption
els used. In case of Langmuir model, the regression coefficients
(R2 ) for the linear equation fittings at different salinities are found Adsorption of the surfactant on sand depends on its dose as
to be greater than 0.950 and at high salinity it is above 0.980 shown in Fig. 9. 1000 ppm concentration of different surfactants
whereas the vales of R2 for the Freundlich isotherm model are (SDS, CTAB, and Tergitol 15-S-7) has been used for adsorption
found to less than 0.950. Therefore, in presence of salt adsorp- study at 303 K. From Fig. 9 it has been found that adsorption
tion of surfactants on sand surface follow the Langmuir isotherm increases with adsorbent dose and then remains constant after
model. certain dose for each surfactant. As the amount of adsorbent

Fig. 8. Adsorption isotherms of SDS at different NaCl salt concentrations at 303 K: (a) Langmuir equation fitting; (b) Freundlich equation fitting.

Table 2
Adsorption isotherm parameters of SDS at different salinities.

Salinity (wt% NaCl) Langmuir parameters Freundlich parameters

 max (mg/g) KL × 10 (L/mg)


2
R2
KF (mg/g) 1/n R2

0 0.763 2.141 0.951 0.231 0.173 0.937


2 1.011 1.472 0.993 0.216 0.218 0.894
4 1.031 1.642 0.988 0.345 0.151 0.899
A. Bera et al. / Applied Surface Science 284 (2013) 87–99 95

3.7. Adsorption kinetics

Adsorption is a physicochemical process that involves the mass


transfer of adsorbate from the liquid phase to the adsorbent surface.
A study of kinetics of adsorption is desirable as it provides infor-
mation about the mechanism of adsorption, which is important to
evaluate efficiency of the process. The experimental data of adsorp-
tion of surfactants on sand particles have been analysed by three
different models viz. Lagergren-first-order equation, second-order
equation and intraparticle diffusion model.

3.7.1. Lagergren-first-order kinetic equation


Lagergren-first-order equation is very well known kinetic equa-
tion. It was first proposed by Lagergren in 1988 to determine the
kinetic process of liquid-solid phase adsorption. The common form
of the equation is

dt
= k1 (e − t ) (7)
dt
On integration of this equation for the boundary condition t = 0
Fig. 10. The effect pH on the adsorption process of the surfactants.
to t = t and  e = 0 to  e =  t , gives:

ln(e − t ) = ln e − k1 t (8)
increases the adsorption sites also increase and the adsorption
process takes place easily with increase in order. After a certain where,  e (mg/g) and  t (mg/g) are the amount of surfac-
adsorbent dose there is no further adsorption because of gathering tants adsorbed on sand particles at equilibrium and at time t (min)
of huge adsorption sites and produced particle interaction among respectively. k1 (min−1 ) is the rate constant of the pseudo-first-
the sand particles in the system. Particle-particle interaction takes order adsorption. The values of k1 can be calculated experimentally
place from high adsorbent concentration which leads to a decrease from the slope of the linear plot of ln( e −  t ) versus t.
in total surface area of the adsorbent and an increase in diffused In Fig. 11(a), adsorption kinetics of SDS surfactant on sand par-
path length [59]. ticle at different concentration at 303 K has been depicted. The
parameters are calculated from the model have been summarized
in Table 3.
3.6. Effect of pH on adsorption of surfactants
3.7.2. Pseudo-second-order kinetic equation
The pH of the aqueous solution is one of the important control- The pseudo-second-order kinetic model equation is expressed
ling parameters in the adsorption of surfactant on reservoir rocks. as follow:
Fig. 10 shows the effect of pH on the extent of adsorption of dif-
ferent surfactants (anionic, cationic and nonionic) on clean sand dt 2
= k2 (e − t ) (9)
surface. The sand particles exhibited high adsorption efficiency at dt
low pH for anionic and nonionic surfactants. As pH increases the and rearranging the Eq. (9) gives
adsorption decreases for anionic surfactant. The adsorption capac-
ity at alkaline solution is lower due to the decrease of positively dt
2
= k2 dt (10)
charged sites on adsorbent and the competition between OH− and (e − t )
anionic surfactant for the adsorption site. A number of research
where, k2 (g/mg/min) is the rate constant of the second-order
works has been reported regarding the effect of pH of solution
equation.
on adsorption of surfactants on rock surfaces [27,60–63]. At low
Now applying the boundary conditions t = 0 to t = t and  e = 0 to
pH, SDS adsorption capacity of sand is high due to acidic nature
 e =  t , the integrated linear form of Eq. (10) can be rearranged to
of the solution which makes the sand surface more positive and
obtain Eq. (11).
that is why the interaction of sand surface with anionic surfactant
SDS is high and hence adsorption capacity is high. In case of Ter- t 1 t
gitol 15-S-7 (nonionic), adsorption decreases up to neutral pH and = + (11)
t k2 e2 e
remains almost constant at alkaline pH region. This can be demon-
strated that the presence of lone pair of electrons of the oxygen The plot of t/ t versus t has been shown in Fig. 11(b). The val-
atom of the ethylene oxide group of ethoxylated nonionic surfac- ues of equilibrium adsorption capacity  e and rate constant k2 ,
tant which is broadly attracted by the positively charged surfaces calculated from the intercept and the slope of the linear plot of
of sand particles at pH values lower than 7. The lower adsorption of t/ t versus t, along with the value of regression coefficient R2 , MSE
the surfactant at alkaline region is due to hydrophobic interaction values are listed in Table 3.
only. As pH of the solution increases adsorption of CTAB surfactant
(cationic) also increases because positively charged head groups of 3.7.3. Intraparticle diffusion model
the cationic surfactant are strongly attracted at high pH with neg- The intraparticle mass transfer diffusion model was proposed
atively charged sand surfaces. So from this study the adsorption of by Weber and Morris [64]. For determination of rate constant and
the surfactants on rock surfaces can be reduced or alter by fixing reaction type, first-order and second-order kinetic models are gen-
the solution pH for nonionic and ionic surfactants which are very erally used. To understand the diffusion mechanism of adsorption
important issue regarding the economic feasibility for surfactant process it is very important to introduce intraparticle diffusion
flooding. model. In this model the fractional approach to the equilibrium
96 A. Bera et al. / Applied Surface Science 284 (2013) 87–99

Fig. 11. The kinetics models for adsorption of SDS surfactant on sand particle at different concentrations at 303 K: (a) Lagergren-first order kinetics; (b) pseudo second order
kinetics; (c) interparticle diffusion kinetics.

0.5 that adsorption of surfactant on sand was a multi-step process;


changes according to a function of (Dt/r 2 ) , where D is the dif-
fusion coefficient within the solid adsorbent and r is the particle involving adsorption on the external surface and diffusion into the
radius. interior [70]. It can be demonstrated from Fig. 11(c) and Table 3
The intraparticle diffusion rate constant can be determined from that other adsorption mechanisms along with diffusion contribute
the following equation [65–68]: in the interactions between the surfactant molecules and sand par-
ticles.
t = Kid t 0.5 + C (12) The high value of R2 and low value MSE obtained from the three
models suggest the applicability of the second-order kinetic model
where, Kid (mg/g/min) is the rate constant of intraparticular dif- to describe the adsorption kinetics data of surfactants onto sand
fusion and C is the intercept. A plot of  t versus t0.5 should be surface and the calculated  e values are in good agreement with
straight line with a slope Kid and intercept C when adsorption mech- the experimental one.
anism follows the intraparticle diffusion process. Ho [69] pointed
out that in case of intraparticle diffusion the  t versus t0.5 plot
must go through the origin and that is sole rate-limiting step. In 3.8. Thermodynamic parameters of adsorption
the present study, no plot passed through the origin. This indicates
that although intraparticle diffusion was involved in the adsorption Both enthalpy and entropy are the key factors to be considered
process, it was not sole rate-controlling step. This also confirms in any process design [71]. The feasibility of the adsorption process

Table 3
Kinetics parameters for the adsorption of surfactant on sand particles at different surfactant concentrations.

Kinetics model Kinetics parameters Surfactant concentration (ppm)

400 800 1000


−1 −2 −2
Lagergren-first-order k1 (min ) 1.249 × 10 1.344 × 10 2.03 × 10−2
 e (mg/g) 0.4667 0.4409 0.4857
R2 0.9657 0.9755 0.9471
MSE 2.6354 2.7661 1.9178

Pseudo-second-order k2 (g/mg/min) 2.501 × 10−2 3.230 × 10−2 4.139 × 10−2


 e (mg/g) 0.810 0.875 0.864
R2 0.991 0.994 0.992
MSE 0.8343 0.6421 0.6678

Intraparticle diffusion kid (mg/g/min) 3.896 × 10−2 3.893 × 10−2 3.318 × 10−2
C 0.1313 0.2319 0.3197
R2 0.988 0.981 0.990
MSE 1.2678 1.1523 1.1235
A. Bera et al. / Applied Surface Science 284 (2013) 87–99 97

Fig. 13. Relationship between Langmuir constant and temperature for SDS adsorp-
Fig. 12. Langmuir equation fitting for adsorption isotherms of SDS at different tem- tion on sand surface.
peratures.

adsorption of surfactant at low temperature attributed to the fact


is clarified by the value of change in Gibbs energy, G◦ (J/mol) and that the adsorption interactions are exothermic in nature. The neg-
it is estimated by applying thermodynamic equation [72,73]: ative value of enthalpy change confirmed the exothermic nature
G◦ = −RT ln KL (13) of the sorption process. Negative value of standard entropy change
confirmed that with increase in temperature the randomness of the
where, R is the universal gas constant (8.314 J/K/mol), T is the tem- molecules at the solid–solution interface decreases during the fix-
perature (K) and KL is the Langmuir constant at temperature T. ation of the surfactant molecules on the active site of sand surfaces.
Again the feasibility and endothermic nature of the adsorption Temperature significantly influences the adsorption of surfac-
process are determined by the entropy change, S◦ (J/mol) and tant on reservoir rock surface. In the present study temperature
enthalpy change, H◦ (J/mol). The dependence of temperature on plays an important role. From Fig. 12, it is clear that with increase
adsorption of surfactant on sand particle was evaluated using van’t in temperature adsorption capacity decreases. Two main impacts
Hoff equation by calculating the values of H◦ and S◦ . of temperature are generally found. Firstly, when temperature
S ◦ H ◦ increases the rate of diffusion of the adsorbate across the exter-
ln KL = − (14) nal boundary layer and interior pores of the reservoir rocks is
R RT
decreased because of the solution viscosity declines as tempera-
G◦ = H ◦ − TS ◦ (15) ture increases. Secondly, temperature influences the equilibrium
adsorption capacity of the sand particles depending on whether
The effect of temperature on adsorption of surfactant on sand the adsorption process is exothermic or endothermic.
particle at different temperature has been depicted in Fig. 12. Tem- Pressure can also play an important role on adsorption of gases
perature plays an important role on the adsorption of surfactant or liquids when physisorption has taken place onto solid surface.
onto sand particles. The variation of adsorption with temperature The amount of adsorption will increase with increase in pressure.
has been explained with help of the thermodynamic parameters The increased adsorption capacity is due to reduction in adsorbate
such as change in standard Gibbs free energy, enthalpy and entropy. volume during adsorption with increase in pressure. It is important
The variation of Langmuir constant with temperature has been to note that the effect of pressure on adsorption of gas is stronger
shown in Fig. 13. The values of S◦ and H◦ were calculated from than liquid on solid surface.
the intercept and slope of plot between lnKL versus 1/T. The cal-
culated values of all the thermodynamic parameters have been
4. Conclusions
reported in Table 4. The negative values of G◦ indicate the sponta-
neous and feasibility nature of surfactant adsorption process. It may
The adsorption of the three types of surfactants namely anionic
also be noted that with increase in temperature from 303 to 333 K,
(SDS), cationic (CTAB), and nonionic (Tergitol 15-S-7) onto clean
the negative values of the Gibbs free energy decrease. This sug-
sand particles from aqueous solutions was systematically stud-
gests that with increase in temperature spontaneity and feasibility
ied. Experimental investigations were carried out to examine the
of the process are decreased and resulting the weaker adsorptive
adsorption equilibrium, isotherm, kinetic behaviors, and thermo-
force. In general, the value of Gibbs free energy for physisorption
dynamics of adsorption of these surfactants. XRD study shows the
lies between −20 kJ/mol and 0 kJ/mol and that for chemisorptions
presence of silica in the pure sand which provides in active sites
lies between −400 kJ/mol and −80 kJ/mol value [74]. The high
for adsorption of different surfactants. FTIR of the sand particles
again indicates the presence of silica. After treatment of surfactant
Table 4 spectral changes are found and adsorption is confirmed from the
Thermodynamic parameters for the adsorption of SDS on clean sand particles at
different temperatures.
result. According to the results obtained in the present study, as we
move from cationic to anionic via nonionic surfactant, adsorption
Temperature (K) KL (L/mol) G◦ (kJ/mol) H◦ (kJ/mol) S◦ (kJ/mol/K) of surfactants on sand particles decreases. With increasing salin-
303 6.113 −4.562 ity of the solution adsorption of SDS increases on sand surface
313 4.959 −4.167 −24.846 −0.0667 due to low electrostatic repulsion between the adsorbed surfactant
323 3.316 −3.217
species. With increase in the surfactant concentration, adsorption
98 A. Bera et al. / Applied Surface Science 284 (2013) 87–99

on the surface of sand particles increases until the saturation point [17] N. Li, G. Zhang, J. Ge, J. Luchao, Z. Jianqiang, D. Baodong, H. Pei, Adsorption
reaches. The sand particles exhibited high adsorption efficiency at behavior of betaine-type surfactant on quartz sand, Energy Fuels 25 (2011)
4430–4437.
low pH for anionic and nonionic surfactants. But opposite trend [18] T. Amirianshoja, R. Junin, A.K. Idris, O.A. Rahmani, Comparative study of sur-
was found for cationic surfactant. Results also show that after a factant adsorption by clay minerals, J. Pet. Sci. Eng. 101 (2013) 21–27.
certain adsorbent dose there is no further adsorption because of [19] S. Zendehboudi, M.A. Ahmadi, A.R. Rajabzadeh, N. Mahinpey, I. Chatzis, Exper-
imental study on adsorption of a new surfactant onto carbonate reservoir
gathering of huge adsorption sites and produced particle interac- samples-application to EOR, The Canadian Journal of Chemical Engineering
tion among the sand particles in the system. Adsorption parameters (2013) 1–11, http://dx.doi.org/10.1002/cjce.21806.
for the Langmuir, Freundlich, Redlich-Peterson, and Sips isotherms [20] M.A. Ahmadi, S.R. Shadizadeh, Experimental investigation of adsorption of a
new nonionic surfactant on carbonate minerals, Fuel 104 (2013) 462–467.
were determined by using experimental data. Results show that the
[21] R. Abdollahi, S.R. Shadizadeh, Experimental investigation of side effect of henna
Langmuir isotherm and pseudo-second order kinetics models suit extract as a new and ecofriendly corrosion inhibitor on acid injectivity of cal-
the equilibrium and kinetics of adsorption on sand surface respec- careous sandstone, J. Transport Porous Media 97 (2013) 105–118.
[22] M.P. Shahri, S.R. Shadizadeh, M. Jamialahmadi, Applicability test of new surfac-
tively. With increasing temperature adsorption of SDS surfactant
tant produced from Zizyphus Spina-Christi Leaves for enhanced oil recovery in
decreases as the randomness of the molecules at the solid-solution carbonate reservoirs, J. Japan Petroleum Inst. 55 (2012) 27–32.
interface decreases during the fixation of the surfactant molecules [23] M. Drach, J. Jabłoński, J. Narkiewicz-Michałek, M. Szymula, Co-adsorption of
on the active site of sand surfaces. surfactants and propyl gallate on the hydrophilic oxide surfaces, Appl. Surf. Sci.
256 (2010) 5444–5448.
[24] A.M. Gaudin, D.W. Fuerstenau, Quartz flotation with cationic collectors, Trans.
AIME 202 (1955) 958–962.
Acknowledgements [25] J.R. Milton, Surfactants and Interfacial Phenomena, third ed., Wiley, New York,
2004.
[26] J.F. Scamehorn, R.S. Schechter, W.H. Wade, Adsorption of surfactants on min-
The authors gratefully acknowledge the financial assistance pro-
eral oxide surfaces from aqueous solutions. Part 1. Isomerically pure anionic
vided by University Grant Commission [F. No. 37-203/2009(SR)], surfactants, J. Colloid Interface Sci. 85 (1982) 463–478.
New Delhi to the Department of Petroleum Engineering, Indian [27] M. Baviere, E. Ruaux, D. Defives, Sulfonate retention by kaolinite at high pH-
School Of Mines, and Dhanbad, India. Thanks are also extended to effect of inorganic anions, SPE Reservoir Eng. 8 (1993) 123–127.
[28] S.G. Dick, D.W. Fuerstenau, T.W. Healy, Adsorption of alkylbenzene sulfonate
all individuals associated with the project. (ABS) surfactants at the alumina–water interface, J. Colloid Interface Sci. 37
(1971) 595–602.
[29] D.W. Fuerstenau, T. Wakamatsu, Effect of pH on the adsorption of sodium
Appendix A. Supplementary data dodecane-sulphonate at the alumina/water interface, Faraday Discuss. Chem.
Soc. 59 (1975) 157–168.
[30] B. Ball, D.W. Fuerstenau, Thermodynamics and adsorption behaviour in the
Supplementary data associated with this article can be quartz/aqueous surfactant system, Discuss. Faraday Soc. 52 (1971) 361–371.
found, in the online version, at http://dx.doi.org/10.1016/ [31] S. Paria, C.K. Kartic, A review on experimental studies of surfactant adsorption
j.apsusc.2013.07.029. at the hydrophilic solid–water interface, Adv. Colloid Interface Sci. 110 (2004)
75–95.
[32] H.W. Sophie, P. Phillip, Adsorption of anionic surfactant by activated carbon:
effect of surface chemistry, ionic strength, and hydrophobicity, J. Colloid Inter-
References face Sci. 243 (2001) 306–315.
[33] M.N. Zhang, X.P. Liao, B. Shi, Adsorption of surfactants on chromium leather
[1] M.J. Lawrence, G.D. Ress, Microemulsion-based media as novel drug delivery waste, J. Soc. Leather Technologists Chemists 90 (2005) 1–6.
system, Adv. Drug Deliv. Rev. 45 (2000) 89–121. [34] R. Atkina, V.S.J. Craigb, E.J. Wanless, S. Biggs, The influence of chain length
[2] C. Czapla, H.J. Bart, Characterization and modeling of the extraction kinetics of and electrolyte on the adsorption kinetics of cationic surfactants at the
organic acids considering boundary layer charge effects, Chem. Eng. Technol. silica–aqueous solution interface, J.Colloid Interface Sci. 266 (2003) 236–244.
23 (2000) 1058–1062. [35] P.A. Siracusa, P. Somasundaran, The role of mineral dissolution in the adsorp-
[3] S.H. Lin, C.M. Lin, H.G. Leu, Operating Characteristics and kinetics studies of tion of dodecylbenzenesulfonate on kaolinite and alumina, Colloids Surf. A 26
surfactant waste water treatment by fenton oxidation, Water Res. 33 (1999) (1987) 55–77.
1735–1741. [36] R. Zhang, P. Somasundaran, Advances in adsorption of surfactant and their mix-
[4] E. Sabah, M. Tarun, M.S. Celik, Adsorption mechanism of cationic surfactants tures at solid/solution interfaces, Adv. Colloid Interface Sci. 123–126 (2006)
onto acid- and heat-activated sepoilities, Water Res. 36 (2002) 3957–3964. 213–229.
[5] W.W. Gale, E.I. Sandvik, Tertiary surfactant flooding: petroleum sulfonate [37] V.S. Singh, B.R. Pandey, Role of adsorption in improved oil recovery by surfac-
composition-efficiency studies, SPE J. 13 (1973) 191–199. tant flooding, J. Sci. Indus. Res. 41 (1982) 54–59.
[6] L.L. Schramm, Surfactants: Fundamentals and Applications in the Petroleum [38] J. Leja, Surface Chemistry of Froth Flotation, Plenum Press, New York, 1982.
Industry, Cambridge University Press, Cambridge, UK, 2000. [39] J. Harkot, B. Jańczuk, The role of adsorption of dodecylethyldimethylammo-
[7] W. Kwok, R.E. Hayes, H.A. Nasr-El-Din, Modelling dynamic adsorption of an nium bromide and benzyldimethyldodecylammonium bromide surfactants in
anionic surfactant on Berea sandstone with radial flow, Chem. Eng. Sci. 50 wetting of polytetrafluoroethylene and poly(methyl methacrylate) surfaces,
(1995) 769–783. Appl. Surf. Sci. 255 (2009) 3623–3628.
[8] F.D.S. Curbelo, V.C. Santanna, E.L. Barros Neto, T.V. Dutra Jr., A.A.A. Castro Dantas [40] X. Wei, X. Wang, J. Liu, D. Sun, B. Yin, X. Wang, Adsorption kinetics of 3-alkoxy-
Neto, I.C. Garnica, Adsorption of nonionic surfactants in sandstones, Colloids 2-hydroxypropyl trimethyl ammonium chloride at oil–water interface, Appl.
Surf. A 293 (2007) 1–4. Surf. Sci. 261 (2012) 237–241.
[9] K.P. Ananthapadmanabhan, P. Somasundaran, Mechanism for adsorption max- [41] W. Stumm, J.J. Morgan, Aquatic Chemistry: An Introduction Emphasizing
imum and hysteresis in a sodium dodecylbenzenesulfonate/kaolinite system, Chemical Equilibria in Natural Waters, J. Wiley and Sons, New York, 1970.
Colloids Surf. A 7 (1983) 105–114. [42] P. Somasundaran, U.L. Huang, Adsorption/aggregation of surfactants and their
[10] A.M. Blokhus, H. Hoiland, M.I. Gjerde, E.K. Erslandb, Adsorption of sodium mixtures at solid liquid interfaces, Adv. Colloid Interface Sci. 88 (2000) 179–208.
dodecyl sulfate on kaolin from different alcohol–water mixtures, J. Colloid [43] P. Somasundaran, D.W. Fuerstenau, Mechanisms of alkyl sulfonate adsorption
Interface Sci. 179 (1996) 625–627. at the alumina-water interface, J. Phys. Chem. 70 (1966) 90–96.
[11] L.K. Koopal, E.M. Lee, M.R. Bohmer, Adsorption of cationic and anionic sur- [44] S. Paria, C. Monohar, K.C. Khilar, Adsorption of anionic and non-ionic surfactants
factants on charged metal oxide surfaces, J. Colloid Interface Sci. 170 (1995) on a cellulosic surface, Colloids Surf. A 252 (2005) 221–229.
85–97. [45] E. Ayranci, O. Duman, Removal of anionic surfactants from aqueous solutions
[12] P. Somasundaran, K.P. Ananthapadmanabhan, K.V. Viswanathan, Adsorption of by adsorption onto high area activated carbon cloth studied by in-situ UV
sulfonate on kaolinite and alumina in the presence of gypsum, in: Soc. Pet. Eng. spectroscopy, J. Hazard. Mater. 148 (2007) 75–82.
AIME, Paper SPE 11780 presented at SPE Oilfield and Geothermal Chemistry [46] M.A. Ahmadi, S. Zendehboudi, A. Shafiei, L. james, Nonionic surfactant for
Symposium, 1-3 June, Denver, Colorado, 1983, pp. 97–104. enhanced oil recovery from carbonates: adsorption kinetics and equilibrium,
[13] P. Somasundaran, S. Krishnakumar, Adsorption of surfactants and polymers at Ind. Engg. Chem. Res. 51 (2012) 9894–9905.
the solid–liquid interface, Colloids Surf. A 123-124 (1997) 491–513. [47] M.A. Ahmadi, S.R. Shadizadeh, Adsorption of novel nonionic surfactant and par-
[14] L.H. Torn, D.A. Keizer, L.K. Koopal, J. Lyklema, Mixed adsorption of poly ticles mixture in carbonates: enhanced oil recovery implication, Energy Fuels
(vinylpyrrolidone) and sodium dodecylbenzenesulfonate on kaolinite, J.Colloid 26 (2012) 4655–4663.
Interface Sci. 260 (2003) 1–8. [48] S.B. Gogoi, Adsorption of non-petroleum base surfactant on reservoir rocks,
[15] K.V. Viswanathan, P. Somasundaran, Adsorption of ethoxylated sulfonates on Curr. Sci. 97 (2009) 1059–1063.
kaolinite and alumina, Colloids Surf. A 26 (1987) 19–41. [49] E. Hoff, B. Nystrom, B. Lindman, Polymer-surfactant interactions in dilute mix-
[16] Z. Xu, X. Yang, Z. Yang, On the Mechanism of Surfactant Adsorption on Solid tures of a nonionic cellulose derivative and an anionic surfactant, Langmuir 17
Surfaces: Free-Energy Investigations, J. Phys. Chem. B 112 (2008) 13802–13811. (2001) 28–34.
A. Bera et al. / Applied Surface Science 284 (2013) 87–99 99

[50] I. Langmuir, The constitution and fundamental properties of solids and liquids: SPE- 147872-MS, presented at SPE Asia Pacific Oil and Gas Conference and
Part I. Solids, J. Am. Chem. Soc. 38 (1916) 2221–2295. Exhibition, Jakarta, Indonesia, 20-22 September, 2011.
[51] M.M. Ayad, A.A. El-Nasr, Adsorption of cationic dye (Methylene Blue) from [62] P. Somasundaran, H.S. Hanna, Adsorption of sulfonates on reservoir rocks, SPE
water using polyanilines nanotubes base, J. Phys. Chem. C 114 (2010) J. 19 (1979) 221–232.
14377–14383. [63] P. Somasundaran, H.S. Hanna, Adsorption/desorption of sulfonate by reservoir
[52] M.A. Wahab, S. Jellali, N. Jedidi, Ammonium biosorption onto sawdust: FTIR rock minerals in solutions of varying sulfonate concentrations, SPE J. 25 (1985)
analysis, kinetics and adsorption isotherms modeling, Bioresour. Technol. 101 343–350.
(2010) 5070–5075. [64] W.J. Weber, J.C. Morriss, Kinetics of adsorption on carbon from solution, J.
[53] B. Samiey, M. Dargahi, Kinetics and thermodynamics of adsorption of Congo Sanitary Eng. Div. Am. Soc. Civ. Eng. 89 (1963) 31–60.
red on cellulose, Central Eur. J. Chem. 8 (2010) 906–912. [65] M.J.D. Low, Kinetics of chemisorptions of gases on solids, Chem. Rev. 60 (1960)
[54] C.B. Vidal, A.L. Barros, C.P. Moura, A.C.A. de Lima, F.S. Dias, L.C.G. Vasconcellos, 267–312.
P.B.A. Fechine, R.F. Nascimento, Adsorption of polycyclic aromatic hydrocar- [66] Y.S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process
bons from aqueous solutions by Modified Periodic Mesoporous Organosilica, J. Biochem. 34 (1999) 451–465.
Colloid Interface Sci. 357 (2011) 466–473. [67] S. Lagergren, Zur theorie der sogenannten adsorption gelöster stoffe, Kungliga
[55] Q.S. Liu, T. Zheng, P. Wang, J.P. Jiang, N. Li, Adsorption isotherm, kinetic and Svenska Vetenskapsakademiens, Handlingar 24 (1898) 1–39.
mechanism studies of some substituted phenols on activated carbon fibres, [68] D.M. Ruthven, Principles of Adsorption and Adsorption Processes, Wiley Inter-
Chem. Eng. J. 157 (2010) 348–356. science, New York, 1982.
[56] T.S. Anirudhan, P. Senan, Adsorption characteristics of cytochrome C onto [69] Y.S. Ho, Removal of cupper ions from aqueous solution by tree fern, Water Res.
cationic Langmuir monolayers of sulfonated poly(glycidylmethacrylate)- 37 (2003) 2322–2330.
grafted cellulose: mass transfer analysis, isotherm modeling and thermody- [70] K.G. Bhattachacharyya, A. Sharma, Kinetics and thermodynamics of methy-
namics, Chem. Eng. J. 168 (2011) 678–690. lene blue adsorption on neem (Azadirachta indica) leaf powder, Dyes Pigm.
[57] S. Chatterjee, M.W. Lee, S.H. Woo, Adsorption of congo red by chitosan hydro- 65 (2005) 51–59.
gel beads impregnated with carbon nanotubes, Bioresour. Technol. 101 (2010) [71] C. Ijagbemi, M. Baek, D. Kim, Montmorillonite surface properties and sorp-
1800–1806. tion characteristics for heavy metal removal from aqueous solutions, J. Hazard.
[58] A.V. Pethkar, K.M. Paknikar, Recovery of gold from solution using Cladospori- Mater. 166 (2009) 538–546.
oides biomass beads, J. Biotech. 63 (1998) 211–220. [72] L. Juang, C. Wang, C. Lee, Adsorption of basic dyes onto MCM-41, Chemosphere
[59] A. Shukla, Y.H. Zhang, P. Dubey, The role of sawdust in the removal of unwanted 64 (2006) 1920–1928.
materials from water, J. Hazard. Mater. 95 (2002) 137–152. [73] M. Wiśniewska, The temperature effect on the adsorption mechanism of poly-
[60] P. Somasundaran, L. Zhang, Adsorption of surfactants on minerals for wett- acrylamide on the silica surface and its stability, Appl. Surf. Sci. 258 (2012)
ability control in improved oil recovery processes, J. Pet. Sci. Eng. 52 (2006) 3094–3101.
198–212. [74] Q.S. Liu, T. Zheng, P. Wang, J.P. Jiang, N. Li, Adsorption isotherm, kinetic and
[61] C.T.Q. Dang, Z. Chen, N.T.B. Nguyen, W. Bae, T.H. Phung, Development of mechanism studies of some substituted phenols on activated carbon fibers,
isotherm polymer/surfactant adsorption models in chemical flooding, in: Paper Chem. Eng. J. 157 (2010) 348–356.

Anda mungkin juga menyukai