Journal of Petroleum Science and Engineering: How Can We Know? Conductivity?

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Journal of Petroleum Science and Engineering 119 (2014) 36–43

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Influences of hydrophilic and hydrophobic silica nanoparticles on


anionic surfactant properties: Interfacial and adsorption behaviors
Mohammad Zargartalebi a, Nasim Barati a,n, Riyaz Kharrat b
a
Petroleum Engineering Department, Petroleum University of Technology, Ahwaz, Iran
b
Tehran Petroleum Research Center, Petroleum University of Technology, Tehran, Iran

art ic l e i nf o a b s t r a c t

Article history: Regarding the novel applications of nanoparticles in enhanced oil recovery, the objective of this study is
Received 17 November 2012 to investigate if nano-sized silica particles have the potential to introduce enhancement in several
Accepted 16 April 2014 aspects of surfactant properties particularly its interfacial and adsorption behaviors. Two types of
Available online 29 April 2014
hydrophilic and slightly hydrophobic fumed silica nanoparticles are used in conjunction with sodium
Keywords: dodecyl sulfate. Extensive series of interfacial tension and adsorption measurement experiments are
silica nanoparticle performed. The results indicate that surfactant interfacial and adsorption properties are interestingly
enhanced oil recovery influenced by the addition of silica particles. Inclusion of both nanoparticles into surfactant solution
surfactant adsorption causes contrasting interfacial behaviors in low and high surfactant concentrations. The adsorption of
interfacial tension
surfactant molecules on the rock surface is generally reduced in the presence of nanoparticles except for
some highly concentrated surfactant solutions.
& 2014 Elsevier B.V. All rights reserved.

how can we know?


1. Introduction conductivity? Despite the considerable surveys investigating different appli-
cations of nanotechnology in reservoir engineering and enhanced
Application of nanotechnology has been very limited in petro- oil recovery techniques, far too little attention has been paid to
leum industry. However, in recent years, the importance of nano- utilization of surfactant/nanoparticle systems as a novel surface-
science to develop conventional methods in several branches of modifier chemical agent in chemical flooding and the studies
petroleum engineering has been highlighted. Lately, several usually subsume interfacial behavior of various types of surfac-
aspects of the likely benefits of novel nano-tech based agent in tants in the presence of nanoparticles. Esmaeilzadeh et al. (2014)
smart fluid design for various oil field applications have been have investigated the influence of ZrO2 on interfacial behavior of
described, particularly for a new generation of drilling (Sensoy et surfactant solutions. According to their results inclusion of nano-
al., 2009), fracturing (Crews and Huang, 2010; Huang and Clark, particles augments the surface activity of negatively charged
2012), completion and stimulation fluids (Crews and Huang, 2008; surfactant, sodium dodecyl sulfate (SDS), below its critical micelle
Huang et al., 2011). concentration (CMC) and as a result it decreases the interfacial
Reservoir engineering, however, has received the most attention tension between oil and water. Ma et al.'s data (Ma et al., 2008),
for nanotechnology applications. Several experiments have been accordingly, presents interfacial tension reduction of anionic
conducted to investigate flow behavior of nanoparticle suspensions surfactant, SDS, in the presence of silica nanoparticles. They
through porous media. Rodriguez et al. (2009) investigated the introduced the electrostatic repulsion between similarly charged
migration of concentrated surface treated silica nanoparticles in particles and surfactant as the main reason for such behavior. The
sedimentary rocks. Kanj et al. (2009) identified the usable size of inclusion of silica nanoparticles in cationic surfactant solution, on
nanoparticles in reservoir rocks through Nano-fluid core flooding the other hand, increases interfacial tension in liquid–air and
experiments. Using hydrophilic/hydrophobic synthesized nanopar- liquid–liquid media (Ravera et al., 2006). This behavior is ascribed
ticles, Zhang et al. (2011) made oil in water emulsions and stabilized to the adsorption of nanoparticles in specific concentrations and
CO2 foams with quite high stability. formation of nanoparticle augmented surfactant layer. This result
is inconsistent with observations of Lan et al. (2007) in which they
studied the effect of silica nanoparticles on cetyltrimethyl ammo-
n
nium bromide's surface activity. Their result has been heralded as
Correspondence to: End of Ayatollah Behbahani Expressway, Petroleum
the lowering impact of nanoparticles on interfacial tension which
Engineering Department, Petroleum University of Technology, P.O. Box 63431,
Ahwaz, Iran. Tel.: 98 9131294402. is attributed to electrostatic attraction between negative nanopar-
E-mail address: nasim.barati@ymail.com (N. Barati). ticles and positive surfactant.

http://dx.doi.org/10.1016/j.petrol.2014.04.010
0920-4105/& 2014 Elsevier B.V. All rights reserved.
M. Zargartalebi et al. / Journal of Petroleum Science and Engineering 119 (2014) 36–43 37

Although the studies conducted on interfacial behavior of 2.1.2. Nanoparticles


surfactants in the presence of nanoparticle are quite valuable, a Two types of nanoparticles, hydrophilic and slightly hydro-
mechanistically engineering approach considering alteration of phobic silica nanomaterials, were used. Both nanoparticles were
dominant factors in surfactant flooding in the presence of nano- supplied by Degussa industries (Evonik). Well-characterized
particle is still missing. fumed Silica, AEROSIL 300 (A300), was the hydrophilic nanopar-
This paper is aimed at studying SDS anionic surfactant solution ticle. These colloidal particles are prepared by hydrolysis of silicon
properties in the presence of hydrophilic and slightly hydrophobic tetrachloride in which silanol groups (Si–OH) are generated on
silica nanoparticles. The main focus is on surfactant interfacial and silica surface. The second nanoparticle used was slightly hydro-
equilibrium adsorption behaviors since a successful chemical phobic type commercially named AEROSIL R816. AEROSIL R816 is
flooding entails having favorable fluid–fluid and fluid–rock beha- fumed silica (AEROSIL 200) after treated with hexadecylsilane
vior. Several hypotheses describing various interface phenomena (C16H33) hydrophobic groups. Table 1 represents some physico-
are used to explain the obtained results. chemical properties of nanoparticles used in this study. It must be
mentioned that both nanoparticles have amorphous structures
and are approximately spherical in shape and their average size is
2. Experimental work of order of few nanometers as observed in TEM images repre-
sented in Fig. 1.
This section contains the information about materials and
experimental procedures used in this study. All the experiments 2.1.3. Rock sample
were performed at ambient conditions (temperature of 30 1C and Sandstone samples of one of Middle East reservoirs were used.
atmospheric pressure). Based on petrographic studies, the rock mainly consisted of quartz,
feldspars and minor amount of accessory minerals like zircon and
2.1. Materials sphene. Sand rocks were initially crushed into small rock fractions
into single grains using a mortar and pestle. The resulting sand
The following materials were used in the experiments. was sieved using meshed sieves under the agitation of a Ro-Tap
Testing Sieve Shaker. The sand grains of appropriate size (400–
2.1.1. Surfactant 500 μm) were gathered and carefully washed with distilled water
The surfactant used in this study was an anionic surfactant and dried in an oven at 150 1C for three hours.
named Sodium Dodecyl Sulfate (SDS) bought from Merck Com-
pany. The CMC value of anionic surfactant was determined to be 2.2. Experiments
approximately 2200 ppm using conductivity measurement tech-
nique. The CMC value was relatively close to those given in Several experiments were conducted which are
literature (Atkin et al., 2003). explained below.

Table 1
Physico-chemical properties of nanoparticles used in this study.

Property Unit Typical value for AEROSIL 300 Typical value for AEROSIL R816

Behavior in the presence of water – Hydrophilic Slightly hydrophobic


Appearance – Fluffy white powder Fluffy white powder
BET surface area m2/g 300 730 1907 20
Average primary particle size nm 7 12
Tamped density gr/l 50 60
PH value – 3.7–4.7 4.0–5.5
C-content wt% – 0.9–1.8
SiO2 wt% Z 99.8 Z 99.8
Al2O3 wt% r 0.050 r 0.050
Fe2O3 wt% r 0.003 r 0.010
TiO2 wt% r 0.030 r 0.030
HCI wt% r 0.025 r 0.025

Fig. 1. TEM images of nanoparticles used; (left) hydrophilic nanoparticle, (right) slightly hydrophobic nanoparticles.
38 M. Zargartalebi et al. / Journal of Petroleum Science and Engineering 119 (2014) 36–43

2.2.1. Surfactant solution preparation 24 h would be enough for adsorption process to reach to equili-
Surfactant solutions were prepared by slow addition of accu- brium conditions.
rate amounts of surfactant powder into distilled water while The surfactant solution and sand sediments were then sepa-
stirring on a magnetic stirrer. The solution was gently heated to rated by centrifugation. Finally, the surfactant concentration was
overcome temperature of solubilizing or Kraft temperature for determined using appropriate calibration curves made by surfac-
30 min. Each solution was prepared separately to eliminate the tant solution conductivity. Each experiment was performed three
error caused by stock solution and dilution. times to eliminate the sources of uncertainty and check the
repeatability of adsorption measurement results.
2.2.2. Nanoparticle dispersion
The method used for nanoparticle dispersion is of great
importance as the solution properties are extremely dependent 3. Results and discussion
on this method. Here, specified amounts of nanoparticles were
first wetted by the aqueous media using a magnetic stirrer. Then, This section covers the experimental work done during this
dispersion and deagglomeration process was done using ultra- study. Results of experiments like stability investigation, interfacial
sonic stirrer for one hour. The ultrasonic device used was a UIP and adsorption measurements are discussed.
1000 hd processor with frequency of 20 KHz and power of
1000 W, manufactured by Hielscher. Ultrasonic cavitation gener- 3.1. Nanoparticle suspension stability
ates high shear that breaks particle agglomerates into single
dispersed particles. Stock solutions of 1 wt% were prepared and In this section, we are mainly concerned with dispersions of
diluted to the desired concentrations. colloidal solid particles in liquids. The main point is that if the
dispersion remains in dispersed state under the experiment
2.2.3. Stability investigations conditions or flocculation and coagulation take place. Two cases
The stability of nanoparticle suspensions was examined by of aqueous suspension stability without any additives and with
visual observation at a particle concentration of 2000 ppm. The surfactant addition were considered.
solutions were observed through high-definition optical equip- While the aqueous suspensions of AEROSIL 300 hydrophilic
ment to see if the dispersion would remain in dispersed state nanoparticles remained stable for more than one month mainly
under the experiment conditions or flocculation and coagulation due to their tiny size, slightly hydrophobic AEROSIL R816 nano-
would take place. The pictures were taken by a computer interface particle suspensions tended to agglomerate after a short period of
attached to the camera. two to three days. This phenomenon may be related to the
hydrophobic chain attached to nanoparticle surface which makes
2.2.4. Interfacial tension measurements the aqueous bulk solution to be an unfavorable environment for
The interfacial tension (IFT) between kerosene and aqueous individual particles. As a result, nanoparticles begin to agglomer-
solutions was determined by axisymmetric drop shape analysis of ate and finally settle due to gravitational forces.
a captive drive instrument (pendant drop method) manufactured Addition of SDS anionic surfactant to nanoparticle suspensions
by Alberta (Canada). The IFT values were obtained using Young– extended the stability period for both nanoparticles especially in
Laplace equation (Bashforth and Addams, 1882). the case of AEROSIL R816 and no agglomeration occurred. Fig. 2
represents both agglomerated and stabilized suspensions for
2.2.5. Adsorption measurements slightly hydrophobic R816 nanoparticles. Flocculated and agglom-
In this study, surfactant adsorption onto prepared sand grains erated particles were observed in the absence of surfactant (left).
was measured using batch technique. To do so, surfactant solu- By the inclusion of surfactant molecules (right), however, the
tions of different concentrations were put in contact with sand suspension remained in stable condition for a long period and
particles with the mass ratio of 5:1 of solution to sand. The no coagulation was observed (The suspension was completely
surfactant solution and sand particles were equilibrated in appro- clear). The stability improvement of nanoparticle suspensions may
priate vessels in gentle horizontal agitation for 24 h. This time is be interpreted through the adsorption of SDS on nanoparticle
more than the time required to reach equilibrium as suggested by surface, bringing about supercharged system by increasing the
Hollander et al. (1981). However, by performing adsorption tests effective charge of silica and consequently stabilize the system
for different contact durations, it was evaluated if this duration i.e. (Ahualli et al., 2011; Luo et al., 2009). In the case of AEROSIL R816

Fig. 2. Slightly hydrophobic nanoparticles, left: agglomeration in aqueous medium after two days, right: no agglomeration in surfactant solution (100 ppm) after 10 days
(clear suspension).
M. Zargartalebi et al. / Journal of Petroleum Science and Engineering 119 (2014) 36–43 39

nanoparticles, hydrophobic interactions between alkyl groups of 2


silica particle and surfactant chain may be also responsible for Region 4
suspension stabilization. 1.6

Region 1

Relative IFT
3.2. Interfacial tension measurements 1.2

Surface tension of surfactant solutions is possibly one of the Region 3


0.8
most common physical properties of such solutions used to
characterize the properties of surfactants in general. Interfacial Region 2
0.4
tension values for SDS surfactant solutions of different concentra-
tions were obtained while addition of nanoparticles.
0
The interfacial effects of AEROSIL 300 nanoparticles when 0 1000 2000 3000 4000 5000 6000 7000
mixed with surfactant solutions are shown in Fig. 3. It can be Surfactant concentration (ppm)
seen that surfactant shows quite strange interfacial behavior in the
Fig. 4. Relative interfacial tensions for aqueous AEROSIL 300 nanoparticle-aug-
presence of nanosilica particles. It seems that the interactions
mented surfactant solutions of different surfactant concentrations at constant
between nanoparticles and surfactant molecules result in two nanoparticle concentrations.
opposing effects in low and high surfactant concentrations. To
more clearly show this phenomenon, some IFT values are tabu-
lated for low, moderate and high surfactant concentrations in concentrations, the interfacial tension value for nanofluid is always
Table 2. below the interfacial tension for surfactant system and continues
Relative interfacial tension values, which are defined as the to decrease to a minimum value at CMC. The reduction in
ratio of the IFT of nanoparticle-augmented surfactant solutions of interfacial tension values was also observed by (Suleimanov
certain nanoparticle concentration to the corresponding surfactant et al. (2011) while addition of light non-ferrous nanoparticles to
solution, are presented in Fig. 4. Nanoparticle concentration was anionic surfactant solutions.
equal to 1000 ppm for all solutions. It can be seen that at very low However, as surfactant concentration reaches to values greater
surfactant concentrations the inclusion of nanoparticles has a than CMC, nanoparticle presence leads to completely different
negligible influence on the oil/water interfacial tension (Region behavior and the oil/water interfacial tension begins to get higher
1). For larger concentrations of surfactants, rather significant values (Region 3). The interfacial tension eventually reaches a
changes occur when particles are present. In the presence of constant value greater than the one obtained by the surfactant
nanoparticles, the interfacial tension drops more steeply at con- itself (Region 4). The interfacial behavior for two last regions can
centrations below the CMC (approximately 2200 ppm), so that the be well illustrated through the final drop shapes of surfactant
interfacial tension is lower for almost all surfactant concentrations solutions in oil medium as represented in Fig. 5. It is visually
when compared to the system without nanoparticles (Region 2). inferred from the pendant drop shapes that after a certain
This fact is consistent with the observations of Esmaeilzadeh et al. surfactant concentration (2000 ppm), the drops begin to grow
(2014) in which they showed that at high enough nanoparticle and larger drops mean higher interfacial tension values. The drop
shape and size are similar for high surfactant concentrations due
20 to constant interfacial tension.
No Nanoparticle The interfacial behavior of surfactant in the presence of silica
AEROSIL 300 (1000 ppm) nanoparticles may be explained in the following way. At zero or
15
very low surfactant concentrations (Region 1), the utilized parti-
IFT (dyne/cm)

cles are completely wetted by water so that they remain in the


10 bulk solution and do not prefer the interface (Ravera et al., 2006;
Ahualli et al., 2011). The behavior observed in region 2 may be
interpreted using different approaches. First is that as the surfac-
5
tant concentration increases, the repulsive electrostatic forces
between the particles and the SDS promotes the surfactant
0
diffusion toward the interface, which leads to a decrease in the
0 1000 2000 3000 4000 5000 6000 7000 surface tension (Ma et al., 2008). Another approach says that
Concentration (ppm) increasing the surfactant concentration implies an increase in the
Fig. 3. Oil/water interfacial tensions for aqueous AEROSIL 300 nanoparticle-
surface coverage of the nanoparticles and, as a consequence, a
augmented surfactant solutions of different surfactant concentrations at constant greater affinity of the particles for the fluid interface (Ravera et al.,
nanoparticle concentrations. 2006). Thus, the number of particles at the interface rises with
surfactant concentration. Such particles containing a large amount
of surfactant can act as carriers of SDS toward the interface. The
Table 2 reduction in interfacial tension could be due to surfactant release
The influence of hydrophilic nanoparticles on interfacial tension of surfactant
from the particles but could also be due to the effect of surfactant-
solutions.
coated particles on the surface tension (Ahualli et al., 2011).
Concentration (ppm) Interfacial tension (dyne/cm) Steeper decrease in interfacial tension values may be directly
related to the nature of nanoparticles and not the nanoparticle–
Surfactant Nanoparticle Surfactant Nanoparticle-augmented
surfactant interactions. The interfacial tension reduction can be
solution surfactant solution
due to the attachment of individual nanoparticles at liquid–liquid
500 1000 7.43 6.10 interface. This attachment usually occurs whether the particles are
1000 1000 3.53 2.61 hydrophobic or hydrophilic (Ghezzi et al., 2001; Schwartz et al.,
2000 1000 2.60 1.81 2001). In addition, according to thermodynamic models based on
4000 1000 2.90 4.26
6000 1000 2.85 4.24
formation of electric double layer and particle interfacial free
energy, the presence of nanoparticles at the liquid interfaces is
40 M. Zargartalebi et al. / Journal of Petroleum Science and Engineering 119 (2014) 36–43

Fig. 5. Final drop shapes for aqueous AEROSIL 300 nanoparticle-augmented surfactant solutions of different surfactant concentrations at constant nanoparticle
concentration.

thermodynamically favorable and results in surface tension reduc- 20


tion (Paunov et al., 2002). In essence, spontaneous adsorption of No Nanoparticle

particles at interface decreases the energy of the system (Paunov AEROSIL R816 (1000 ppm)
15 AEROSIL R816 (2000 ppm)
et al., 2002). The adsorption of the particles also changes the
IFT (dyne/cm)

entropy of the dispersions. As the colloidal particles move from


the bulk fluid to the interface, this will create more free volume for 10
the solvent molecules and therefore the entropy of the whole
system should increase (Dong and Johnson, 2003). The combined
decrease of the total internal energy and the increase in the 5
total entropy cause the total free energy of the system to decrease.
Based on thermodynamic definition of the interfacial energy (Defay
0
and Prigoin, 1966), this effect causes the interfacial tension to 0 1000 2000 3000 4000 5000 6000 7000
decrease. Concentration (ppm)
As surfactant concentration increases, greater numbers of
Fig. 6. Oil/water interfacial tensions for aqueous R816 nanoparticle-augmented
nanoparticles are brought to the interface. Therefore, the increase
surfactant solutions of different surfactant concentrations at constant nanoparticle
in interfacial tension value in region 3 may be described using concentrations.
capillary force concept (Dong and Johnson, 2003). When particles
are placed at an interface, they either freely float at the interface or
partially immerse into a liquid layer. For both conditions, lateral of the colloidal dispersion will not change with an increase in the
capillary forces are generated. At high particle concentrations concentration (Dong and Johnson, 2003).
when the particles are close to each other, the attractive capillary The same trend was observed for the case of AEROSIL R816 and
force is quite high and almost uniform across the interface. The SDS mixtures as shown in Fig. 6. However, various interfacial
high capillary force resists the interface deformation and the tension values were observed for the solutions of equal surfactant
interfacial tension is increased. but different nanoparticle concentrations as tabulated in Table 3
Finally, some theories are presented to explain the interfacial for some solutions. When nanoparticle concentration is increased
phenomena occurred in region 4. The first approach suggests that from 1000 to 2000 ppm, oil/water IFT values are higher indepen-
the charged nanoparticles attract the surfactants to aggregate dent of the interfacial region. This may be due to the high affinity
around the nanoparticles and unevenly distribute at the water– of partially hydrophobic particles to remain at the oil–water
oil interfaces (Luo et al., 2009). The interactions between charged interface due to their amphoteric characteristics (Zargartalebi et
nanoparticles and surfactant carbon chains and the electrostatic al., 2013). The surfactant molecules are consequently unable to
interactions between sodium ions, charged particles and surfac- detach the particles from the interface and higher IFT values are
tant head groups alter the surfactant molecules to a relatively mainly due to the smaller number of surfactant molecules present
preferable bended orientation. The interfacial tension values con- at the interface. Nonetheless, the interfacial tension values in
sequently stabilize at higher values respect to surfactant-only region 2 are still lower than the surfactant-only solutions due to
solution. The second hypothesis based on capillary force concept similar mechanisms explained before in the case of hydrophilic
states that when the adsorption of the particles at the interface nanoparticles. Similar trend was observed by Vashisth et al. (2010)
reaches a saturated state, the capillary forces will no longer where they claimed that below the CMC, combination of surfac-
increase. Therefore, at higher concentrations, the surface tension tant and nanoparticles could work like mixed surfactant solutions.
M. Zargartalebi et al. / Journal of Petroleum Science and Engineering 119 (2014) 36–43 41

Table 3 10
The influence of slightly hydrophobic nanoparticles on interfacial tension of SDS

Adsorption (mg SDS/gr Rock)


surfactant solutions. SDS + AEROSIL 300
SDS + AEROSIL R816
Concentration (ppm) Interfacial tension (dyne/cm)

Surfactant Nanoparticle Surfactant Nanoparticle-augmented


1
solution surfactant solution

500 1000 7.43 3.71


500 2000 7.43 4.67
1000 1000 3.53 2.59
1000 2000 3.53 2.76
2000 1000 2.60 1.87 0.1
2000 2000 2.60 2.42 100 1000 10000
4000 1000 2.90 3.64 Surfactant concentration (ppm)
4000 2000 2.90 4.74
6000 1000 2.85 4.26 Fig. 7. Surfactant adsorption behavior in the presence of hydrophilic and slightly
6000 2000 2.85 4.64 hydrophobic nanoparticles.

3.3.1. Ordinary rock/nanoparticle-augmented surfactant solution


The effect of AEROSIL 300 hydrophilic nanoparticles with
Table 4 constant concentration of 1000 ppm on the surfactant adsorption
Negligible influence of nanoparticles on surfactant solution conductivity. is shown in Fig. 7. It can be seen that the amount of surfactant
adsorption has been reduced for a wide range of surfactant
Surfactant Conductivity (μS/s)
concentrations from low concentrations up to a specific concen-
concentration (ppm)
No AEROSIL 300 AEROSIL R816
tration around CMC. After that, higher surfactant adsorption
nanoparticle (1000 ppm) (1000 ppm) respect to sole surfactant solution is observed. The strange
behavior is that surfactant adsorption continues to grow even
100 102 103 105 after critical micelle concentration. Finally, the adsorption amount
250 135 137 140
reaches a constant value greater than the adsorption for sole
500 218 221 220
1000 417 420 422 surfactant solution at higher surfactant concentrations.
1500 562 566 567 This behavior may be related to the interactions between
2000 686 689 688 surfactant molecules and similarly charged individual particles.
3000 924 927 925
Ionic surfactant adsorption is particularly sensitive to the interac-
4000 1060 1062 1060
6000 1353 1357 1360
tions of counter and co-ions with the charged groups of the rock
surface. These can include the level of dissociation of surface
groups and the overall ionic strength. If the affinity of co-ions for
In region 3, a considerably sharper increase in interfacial tension is surface groups is sufficiently high, then the co-ions can compete
observed by increase in nanoparticle concentration. The invasion for adsorption sites at the surface (Paria and Khilar, 2004). In other
of nanoparticles toward the interface is intensified when nano- words, there exists a competition between anionic surfactant
particle concentration gets higher. So, the capillary forces which molecules and negatively charged nanoparticles to be adsorbed
are considered to be the main reason for the increase in interfacial on the rock surface. At low surfactant concentrations, both
tension come into play faster and more severely. As a result, the materials are adsorbed and the number of adsorbed surfactant
sudden increase in IFT would occur. The interfacial tension finally molecules is smaller respect to sole surfactant solution case (lower
reaches a value slightly higher than the previous one obtained adsorption amount). As the surfactant concentration increases, the
with lower nanoparticle concentration. This is due to the fact that tendency of surfactant molecules to be adsorbed gets higher and
with larger number of nanoparticles in the interface region, there higher until the surfactant molecules almost completely win the
exist less volume for the surfactant molecules and surfactant competition and only small amounts of nanoparticles are adsorbed
efficiency in IFT reduction is restricted. on the adsorbent surface.
In the second case, the effect of AEROSIL R816 slightly hydro-
phobic nanoparticles on the surfactant adsorption behavior was
3.3. Adsorption measurements investigated. The nanoparticle concentration was similar to that of
hydrophilic one in the former case. It was seen that the amount of
This section targets the effect of hydrophilic and slightly adsorbed surfactant was smaller for all surfactant concentrations
hydrophobic silica nanoparticle presence on the adsorption beha- as shown in Fig. 7. It is observed that slightly hydrophobic
vior of SDS anionic surfactant. Regarding the fact that the presence nanoparticles have been more successful to reduce the amount
of nanoparticles has negligible influence on surfactant solution of surfactant adsorption respect to hydrophilic ones.
conductivity (Table 4), this property was used to determine It may be interpreted through the interactions between nano-
surfactant adsorption on the rock surface. particles and surfactant molecules. First, the hydrophobic nano-
Several cases were considered and the adsorption behaviors particles are in very close competition with surfactants even at
were compared respect to each other. In the first case, hydrophilic/ high surfactant concentrations. On the contrary, hydrophilic nano-
hydrophobic nanoparticle-augmented surfactant solutions of con- particles are losers when the surfactant concentration is high. The
stant nanoparticle concentration were put in contact with closer competition can be ascribed to the attached hexadecylsilane
crushed rock. groups which render the structure of both surfactant and R816
In the next cases, surfactant adsorption on nanoparticle coated comparable. In other words, the amphoteric structure of R816
rock was investigated. Coated rock was obtained by putting it in nanoparticles augments their affinity to move toward solid–liquid
contact with hydrophilic nanoparticle suspensions of specified interface. As a result, the slightly hydrophobic nanoparticles are
concentration for at least 24 h. more efficient in surfactant adsorption reduction.
42 M. Zargartalebi et al. / Journal of Petroleum Science and Engineering 119 (2014) 36–43

10 mechanisms responsible for surfactant adsorption. Nanoparticle


desorption hypothesis as well as detachment mechanism seem
Adsorption (mg SDS/gr Rock)

to be reasonable since they make the systems alike. The lower


adsorption in low surfactant concentrations however, shows
1
that the detachment process is too slow and incomplete in this
region.

0.1
SDS
3.3.3. Nanoparticle coated rock/nanoparticle-augmented surfactant
SDS + AEROSIL 300
SDS (*) solution
SDS + AEROSIL 300 (*) Finally, in the last case, the adsorption behavior of nanoparticle-
0.01 augmented surfactant solutions on surface coated rock was investi-
100 1000 10000
gated. The nanoparticle concentration was the same in rock surface
Surfactant concentration (ppm)
coating process and augmented surfactant solution preparation (i.e.
Fig. 8. Surfactant and nanoparticle-augmented surfactant adsorption on nanopar- 1000 ppm). The adsorption results are presented in Fig. 8.
ticle surface coated rock in comparison with its adsorption on ordinary rock. It is seen that the adsorption amount is considerably lower for all
(*) denotes that the rock surface has been coated by hydrophilic nanoparticles
surfactant concentrations respect to the situation in which sole
surfactant molecules are to be adsorbed on ordinary rock surface.
Secondly, the presence of slightly hydrophobic nanoparticles This may be explained through the limited surface area available for
retains the surfactant molecules in the bulk solution and prevents surfactant molecules which are also in competition with suspended
their adsorption on the rock surface. This may be attributed to the nanoparticles in the bulk solution. It must be mentioned that as
negatively charged clusters formed by the nanoparticles and nanoparticles exist in the bulk solution, nanoparticle desorption does
surfactant molecules which tend to remain in the bulk solution not occur.
rather than adsorption on the solid due to the electrostatic A comparison between the cases in which surfactant molecules
repulsion between the clusters and charged solid surface. The of surfactant solutions and nanoparticle-augmented surfactant
formation of such charged aggregates was also reported by solutions are to be adsorbed on the surface coated rock is made
Esmaeilzadeh et al. (2012) where the adsorption of surfactant/ in Fig. 8. It can be easily seen that the adsorption amount is less for
nanoparticle mixture on the positively charged carbonate rock nanoparticle-augmented surfactant solutions of all surfactant
surface was increased due to the electrostatic attraction between concentrations.
the aggregates and the solid surface. The increase in surfactant The difference between the curves becomes more and more
adsorption on the carbonate surface has been also reported by obvious at higher surfactant concentrations. At low surfactant
Suleimanov et al. (2011). concentrations both curves show almost similar adsorption beha-
vior. The small difference is due to the fact that for the case of sole
surfactant solution, both desorption and the detachment mechan-
3.3.2. Nanoparticle coated rock/surfactant solution isms of nanoparticles provides wider available surface for surfac-
The effect of surface coated rock on the surfactant adsorption tant molecules. As for the case of nanoparticle augmented
behavior of surfactant solutions is presented in Fig. 8. It can be surfactant solution, however, the lack of nanoparticle diffusion
seen that considerable reduction has occurred in adsorption from solid surface to the bulk solution renders the surfactant
amount for low surfactant concentrations. However, this effect is adsorption to be lower. On the other hand, as the concentration of
gradually diminished at higher surfactant concentrations. The surfactant increases the detachment mechanism becomes more
adsorption amount reaches a plateau higher than the case in significant for the case of nanoparticle free solution and as such
which ordinary rock i.e. non surface coated rock has been used. the amount of surfactant adsorption increases. This mechanism is
Generally, when rock surface is coated with nanoparticles, the weakened for the nanoparticle augmented surfactant solution due
surface available for surfactant molecules and consequently sur- to retention of surfactant in the bulk solution by nanoparticles as
factant adsorption is reduced. Although surfactant may be described before.
adsorbed on nanoparticle surface, it is quite small in comparison
with surfactant adsorption on the rock. However, the results
implicate another mechanism in the adsorption process which Table 5
causes the high adsorption amount at high surfactant concentra- The influence of nanoparticles on surfactant adsorption in different solutions.
tions. This phenomenon may be explained through nanoparticle
Solution Surfactant concentration Surfactant adsorption
desorption. When nanoparticle-free surfactant solution is put in (ppm) (mg/g)
contact with nanoparticle coated rock, the concentration gradient
formed may cause nanoparticle desorption through the bulk SDS 300 0.74
solution. In other words, concentration gradient activates nano- SDS (n) 300 0.10
SDS þ AEROSIL 300 300 0.31
particle diffusion according to the Fick's law. Consequently, the SDS þ AEROSIL 300 (n) 300 0.10
competition between surfactant molecules and nanoparticles SDS þ AEROSIL R816 300 0.12
begins as discussed before. Another mechanism describing such SDS 2000 1.83
adsorption behavior lies in a virtually inevitable fact that nano- SDS (n) 2000 0.62
SDS þ AEROSIL 300 2000 0.58
particles can be detached from the solid surface by surfactant
SDS þ AEROSIL 300 (n) 2000 0.31
molecules. This behavior is just like the particle detachment SDS þ AEROSIL R816 2000 0.19
process at liquid–liquid interface proposed by Vashisth et al. SDS 4000 1.90
(2010). SDS (n) 4000 2.65
Fig. 8 shows that surfactant adsorption behavior at high SDS þ AEROSIL 300 4000 2.70
SDS þ AEROSIL 300 (n) 4000 0.61
surfactant concentrations is similar for nanoparticle-augmented SDS þ AEROSIL R816 4000 1.05
surfactant solution/ordinary crushed rock and surfactant solution/
n
surface coated crushed rock systems. This behavior shows similar Denotes that the rock surface has been coated by hydrophilic nanoparticles.
M. Zargartalebi et al. / Journal of Petroleum Science and Engineering 119 (2014) 36–43 43

The surfactant adsorption values at different surfactant con- Esmaeilzadeh, P., Fakhroueian, Z., Bahramian, A., Arya, S., 2012. Influence of ZrO2
centrations are listed in Table 5 for various scenarios. nanoparticles including SDS and CTAB surfactants assembly on the interfacial
properties of liquid–liquid, liquid–air and liquid–solid surface layers. J. Nano
Res. 21, 15–21.
Esmaeilzadeh, P., Hosseinpour, N., Bahramian, A., Fakhroueian, Z., Arya, S., 2014.
4. Conclusions Effect of ZrO2 nanoparticles on the interfacial behavior of surfactant solutions
at air–water and n-heptane–water interfaces. Fluid Phase Equilib. 361,
289–295.
The influences of hydrophilic and slightly hydrophobic silica Ghezzi, F., Earnshaw, J.C., Finnis, M., McCluney, M.J., 2001. Pattern formation in
nanoparticles on SDS anionic surfactant properties were investi- colloidal monolayers at the air–water interface. J. Colloid Interface Sci. 238 (2),
gated for enhancing oil recovery purpose. Surfactant interfacial 433–446.
Hollander, A.F., Somasundaran, P., Gryte, C.C., 1981. Adsorption from Aqueous
and adsorption behaviors were investigated in the presence of Solutions. Plenum Press, New York.
nanoparticles. Considerably stable nanoparticle suspensions were Huang, T., Clark, D.E., 2012. Advanced Fluid Technologies for Tight Gas Reservoir
obtained adding low amounts of SDS. Stimulation, SPE Saudi Arabia Section Technical Symposium and Exhibition.
Society of Petroleum Engineers, Al-Khobar, Saudi Arabia.
Opposing interfacial properties were observed for low to high Huang, T., Clark, D.E., Crews, J.B., 2011. Protecting the reservoir with surfactant
surfactant concentrations for both hydrophilic and hydrophobic micellar drill-in fluids in carbonate-containing formations. SPE Drill. Complet.
nanoparticle-augmented solutions. The inclusion of nanoparticles 26 (4), 492–498.
Kanj, M.Y., Funk, J.J., Al-Yousif, Z., 2009. Nanofluid Coreflood Experiments in the
had a negligible influence on the oil/water interfacial tension at ARAB-D, SPE Saudi Arabia Section Technical Symposium and Exhibition,
very low surfactant concentrations. However, surfactant efficiency AlKhobar, Saudi Arabia.
was rather significantly improved at higher SDS concentrations up Lan, Q., et al., 2007. Synergistic effect of silica nanoparticle and cetyltrimethyl
ammonium bromide on the stabilization of O/W emulsions. Colloids Surf. A:
to about CMC after which the oil/water interfacial tension began to
Physicochem. Eng. Asp. 302, 126–135.
grow. The interfacial tension eventually reached a constant value Luo, M., Song, Y., Dai, L.L., 2009. Heterogeneous or competitive self-assembly of
greater than the one obtained by the sole surfactant solution of the surfactants and nanoparticles at liquid–liquid interfaces. Mol. Simul. 35 (10–11),
same concentration. 773–784.
Ma, H., Luo, M., Dai, L., 2008. Influences of surfactant and nanoparticle assembly on
From the adsorption point of view, nanoparticles could gen- effective interfacial tensions. Phys. Chem. Chem. Phys 10, 2207–2213.
erally lower the surfactant amount adsorbed on the rock surface. Paria, S., Khilar, K.C., 2004. A review on experimental studies of surfactant
This reduction was much more considerable for hydrophobic adsorption at the hydrophilic solid–water interface. Adv. Colloid Interface Sci.
110 (3), 75–95.
particles in all surfactant concentrations. By the use of nanopar- Paunov, V.N., Binks, B.P., Ashby, N.P., 2002. Adsorption of charged colloid particles
ticle coated rocks, surfactant adsorption was minimized particu- to charged liquid surfaces. Langmuir 18 (18), 6946–6955.
larly at high surfactant concentrations. Ravera, F., Santini, E., Loglio, G., Ferrari, M., Liggieri, L.J., 2006. Effect of nanoparticles
on the interfacial properties of liquid/liquid and liquid/air surface layers. Phys.
Chem. B 110 (39), 19543–19551.
References Rodriguez, E., Roberts, M.R., Yu, H., Huh, H., Bryant, S.L., 2009. Enhanced migration
of surface-treated nanoparticles in sedimentary rocks, in: SPE Annual Technical
Conference and Exhibition, New Orleans, Louisiana, USA.
Ahualli, S., et al., 2011. Adsorption of anionic and cationic surfactants on anionic Schwartz, H., Harel, Y., Efrima, S., 2001. Surface behavior and buckling of silver
colloids: supercharging and destabilization. Langmuir 27, 9182–9192. interfacial colloid films. Langmuir 17 (13), 3884–3892.
Atkin, R., Craig, V.S.J., Wanless, E.J., Biggs, S., 2003. Mechanism of cationic surfactant Sensoy, T., Chenevert, M.E., Sharma, M.M., 2009. Minimizing water invasion in shale
adsorption at the solid-aqueous interface. Adv. Colloid Interface Sci. 103, using nanoparticles, in: SPE Annual Technical Conference and Exhibition, New
219–304. Orleans, LA, USA.
Bashforth, S., Addams, J.C., 1882. An Attempt to Test the Theory of Capillary Action. Suleimanov, B.A., Ismailov, F.S., Veliyev, E.F., 2011. Nanofluid for enhanced oil
Cambridge University Press and Deighton, Bell and Co, London. recovery. J. Pet. Sci. Eng. 78, 431–437.
Crews, J.B., Huang, T., 2008. Performance Enhancements of Viscoelastic Surfactant Vashisth, C., Whitby, C.P., Fornasiero, D., Ralston, J., 2010. Interfacial displacement of
Stimulation Fluids With Nanoparticles, Europec/EAGE Conference and Exhibi- nanoparticles by surfactant molecules in emulsions. J. Colloid Interface Sci. 349,
tion, Rome, Italy. 537–543.
Crews, J.B., Huang, T., 2010. New Remediation Technology Enables Removal of Zargartalebi, M., Kharrat, R., Barati, N., Zargartalebi, A., 2013. Slightly hydrophobic
Residual Polymer in Hydraulic Fractures, SPE Annual Technical Conference and silica nanoparticles for enhanced oil recovery: interfacial and rheological
Exhibition. Society of Petroleum Engineers, Florence, Italy. behavior international journal of oil. Gas Coal Technol. 6 (4), 408–421.
Defay, R., Prigoin, I., 1966. Surface Tension and Adsorption. Longmans, London. Zhang, T. et al., 2011. Engineered Nanoparticles as Harsh-Condition Emulsion and
Dong, L., Johnson, D., 2003. Surface tension of charge-stabilized colloidal suspen- Foam Stabilizers and as Novel Sensors, Offshore Technology Conference,
sions at the water–air interface. Langmuir 19, 10205–10209. Houston, Texas, USA.

You might also like