Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Colloids and Surfaces A 556 (2018) 253–265

Contents lists available at ScienceDirect

Colloids and Surfaces A


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

Application of functionalized silica-graphene nanohybrid for the enhanced T


oil recovery performance

Sanaz Tajika, Abbas Shahrabadia, Alimorad Rashidia, , Milad Jaliliana, Amir Yadegarib
a
Nanotechnology Research Center, Research Institute of Petroleum Industry (RIPI), P.O. Box 14665-137, Tehran, Iran
b
Department of Developmental Sciences, Marquette University School of Dentistry, Milwaukee, WI, 53233, USA

G R A P H I C A L A B S T R A C T

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

Keywords: Nanoparticles as a part of nanotechnology have drawn numerous attentions for its great potential of enhancing oil
Functionalized silica-graphene nanohybrids recovery. However, the role of nanoparticles in this field is still in its infancy and consequently has attracted much
Interfacial tension more attention in the last decade. The main objective of the current study is the experimentally investigation of
Micromodel functionalized silica-graphene nanohybrids as a novel stabilizer of water-decalin and water-asphaltene emulsions
Pickering emulsion
and its effects on enhanced oil recovery (EOR). The conventional properties of crude oil and the prepared na-
nofluids were measured to examine the EOR mechanisms. The nanofluids were prepared by synthetic brine (4 wt%
NaCl) as the base fluid at various concentrations of the functionalized nanohybrids including 0.01. 0.05, and
0.1 wt%. The phase behavior of the nanohybrids dispersions in presence of NaCl was visually observed. The results
showed that the IFT between oil and water can be reduced to about half of its initial value by employing func-
tionalized silica-graphene nanohybrids. The values of IFT decrease via increasing the concentration of nanofluids
which indicates a great potential application of the proposed nanofluid for EOR. Moreover, a visualization flooding
method (glass micromodel) was employed as a porous media to investigate the EOR mechanisms of the nanofluids.

1. Introduction oil can be recovered from the preliminary extraction steps, enhanced oil
recovery (EOR) has been widely employed for more utilization of oil
The depletion of fossil fuels has recently rung an alarming bell for reserves [1–3]. Up to now, various methods and techniques such as
more exploitation from oilfield. Due to the fact that only 50% of crude water flooding, heat treatment, injection of polymers and surfactants


Corresponding author.
E-mail addresses: sanaz_tajik@aut.ac.ir (S. Tajik), rashidiam@ripi.ir (A. Rashidi).

https://doi.org/10.1016/j.colsurfa.2018.08.029
Received 16 June 2018; Received in revised form 7 August 2018; Accepted 9 August 2018
Available online 11 August 2018
0927-7757/ © 2018 Elsevier B.V. All rights reserved.
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

have been investigated as the most prominent approaches for EOR to the direct functionalization. Practically, the proposed approach
[4–7]. Although some of them still suffer from disadvantages such as eliminates the additional steps such as separating, washing, neu-
sophisticated instruments and low productivity. Hence, developing tralizing, purifying, and drying compared to other reported functiona-
novel and applicable techniques for improving the EOR performance is lization methods. To find the best level of functionalization, the time of
in great interest. It has been widely demonstrated that after initial oil exposure and proportion of sulfuric and nitric acid were thoroughly
extraction, water flooding can be applied as an efficient and economic examined and optimized. The as-prepared functionalized nanohybrid
technique for keeping the downhole pressure [8]. However, considering was used as a novel additive for reducing the IFT values of the crude oil-
the viscosity and displacement of the crude oil, waterflooding has its brine (4 wt% NaCl). Furthermore, the stability of oil in water emulsions
own drawbacks. Therefore, developing an appropriate alternative as the by adding the functionalized nanohybrid was studied by taking mi-
tertiary enhanced oil recovery is still demanding. Throughout the last croscopy images for evaluating the average droplet diameters. Even-
decade, employing of nanofluids instead of conventional fluids led to tually, the behavior of the prepared nanofluid in porous media was
high performance oil and gas recovery due to their unique properties visually characterized by designing a glass micromodel.
[9,10]. For instance, economically speaking, addition of only 0.01 wt%
nanoparticles could significantly accelerate the EOR process. The pre- 2. Experimental
sence of nanoparticles may considerably facilitate the EOR process by
enhancing the well drilling, changing the wettability of rocks, boosting 2.1. Materials
the mobility of the trapped oil and decreasing the interfacial tension
[11,12]. It should be noted that the impact of nanoparticles on the EOR All the chemical reagents including sulfuric acid (H2SO4, 98%),
performance and oil displacement mechanism need to be further ex- nitric acid (HNO3, 65%), hydrochloric acid (HCl 37%), citric acid-hy-
plored. The alteration of wettability is considered as a critical para- drate, sodium chloride, acetone, toluene, ethylene chloride, methanol,
meter which should be thoroughly investigated during EOR process trichloromethylsilane, decahydronaphthalene (decalin, > 98%), to-
[13,14]. Moreover, the estimation of IFT values between the aqueous luene (99.9%) and water-glass (Na2SiO3-specific gravity 1.05) were
and oil phases during nanofluid flooding is another effective parameter. purchased from Merck. Deionized water (18 MΩ.cm) was used
It has been well documented that IFT can affect the mobilization of throughout the experiments. The crude oil was provided from one of
trapped oil through capillary forces [15,16]. Although, in the case of the current oil field in Iran and asphaltene separated from this crude oil.
complex nanofluids, comprehensive and experimental studies for cal- The API specifications of the crude oil including specific gravity, visc-
culation of IFT are very rare, there are a few reports that indicate the osity, and density at ambient temperature were estimated to be 40.5
simultaneous using of nanoparticles and anionic surfactant would API, 0.64 cP, and 0.84 g/ml, respectively.
dramatically reduce the IFT values. Among the conventional nanoma-
terials, SiO2 nanoparticles have shown high compatibility for improving 2.2. Synthesis of silica-graphene nanohybrids and their functionalization
the wettability of the rocks and decreasing the IFT [17]. Hendraningrat
et al have used hydrophilic SiO2 nanoparticles as an effective additive The silica-graphene nanohybrid was synthesized according to our
for improving oil recovery [18]. Their experimental studies showed that previous report [30]. Briefly, the silica aerogel powder was prepared
the concentration of nanoparticles should be optimized to get the through a sol-gel method [31]. The obtained powder was annealed at
highest oil recovery in low permeable reservoirs [19]. However, it has 600 °C (5 °C/min) in a horizontal quartz tube furnace and the presence
been suggested that the combination of SiO2 nanoparticles with Fe2O3, of hydrogen flow (600 sccm). After that, acetylene was used as the
MgO, and Al2O3 will provide better results compared to bare silica carbon precursor with the flow rate of 100 sccm for 30 min. Finally, the
nanoparticles [20]. Various kind of organic and inorganic materials reactor was cooled down under nitrogen flow to room temperature.
such as solid lipid, silica, carbon, titanium dioxide, and ferric oxide The functionalization of silica-graphene nanohybrid was carried out
nanoparticles have been utilized for stabilization of oil-water Pickering in vapor phase by a simple setup (see Fig. S1) as described by our
emulsions [21–24]. A small amount of these nanoparticles can produce previous report [32]. It should be noted the silica-graphene nanohybrid
stable emulsions with suitable size distribution and superior stability. was successfully functionalized in this procedure without direct contact
Indeed, the addition of nanoparticles replaces the liquid-liquid inter- of the nanohybrid with liquid phase. The functionalization process was
faces with solid-liquid interfaces which possess less interfacial energy conducted at different time period including 24, 48, and 72 h and the
[25]. Obviously, nanoparticles would assist the dispersion of im- obtained products are denoted as fsg.24, fsg.48, and fsg.72, respec-
miscible droplets into the other liquid leading to produce facile, stable, tively. Fig. 1 shows the scheme of the functionalization of the silica-
and cost-effective emulsifiers [26,27]. The visualization of fluid’s flow graphene nanohybrid surface.
can be investigated by using a two-dimensional porous micromodel
patterns which geometrically simulate a real rock. Generally, micro- 2.3. Preparation of Pickering emulsion stabilized by functionalized silica-
models provide an opportunity for a realistic observation of porous graphene nanohybrids
media in oil recovery which is not possible by other techniques such as
corefloods. However, employing micromodels has been restricted due All of the Pickering emulsions were prepared by deionized water
to their lack of heterogeneous surface on which flow dynamics can be and synthetic brine (4 wt% NaCl) as the aqueous phase and decalin and
dictated. Buchgraber and coworkers designed a two-dimensions mi- asphaltene in toluene as the organic phase with different ratio. The
cromodels on microfluidic devices with an analogous porosity to real optimization suggested that the ratio of (1:1 v: v) showed the higher
rocks for the simulation of porous media [28,29]. In the present study, a stability among the other proportions. In a typical water-oil emulsion,
nanohybrid containing silica nanoparticles and graphene was synthe- 0.1 wt% of the functinalized silica-graphene nanohybrid was ultra-
sized through chemical vapor deposition method. The main advantage sonically dispersed in aqueous phase for 30 min by using ultrasonic
of this procedure is related to the usage of silica nanoparticles as the bath at 100 W. Subsequently, decalin (or various concentrations of as-
heterogenous catalyst which eliminates further purification and se- phaltene in toluene) was added and the mixture was further sonicated
paration of unreacted nanoparticles. Indeed, the silica nanoparticles at 100 W for 15 min. Finally, the obtained Pickering emulsion was
have been simultaneously used as the catalyst for growing the graphene gently dropped on a glass slide for taking microscopic images.
sheets and main part of the nanohybrid. Consequently, the obtained
nanohybrid was chemically functionalized by the evaporation of sul- 2.4. Interfacial tension (IFT) measurement
furic and nitric acid in gas phase. This technique noticeably promotes
the preparation of the functionalized silica-graphene nanohybrid owing For measuring the IFT values between crude oil and nanofluids the

254
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

Fig. 1. Scheme of the functionalization of the silica-graphene nanohybrid through treatment with a mixture of nitric and sulfuric acid vapors.

Table 1 The procedure was followed by the injection and circulation of pure
Micromodel physical properties. acetone to the micromodel. At the final step, the preparation procedure
Length 6 cm
was finished by washing the micromodel with deionized water. The
Width 4.5cm treated micromodel can be used for modeling oil-wet behavior through
Depth 0.1mm a porous medium. In a typical experiment, crude oil was firstly injected
Porosity 36.97% to prepare an oil-saturated micromodel. Then, the NaCl solution (4 wt
Pore Volume 0.1 cc
%) was constantly injected into the oil-saturated micromodel at the rate
Absolute permeability 11 md
of 0.03 ml/hr till no more oil was produced. Thereafter, the as prepared
nanofluids were injected with the flow rate of 0.03 ml/hr as long as the
DSA technique was employed. The DSA technique has been introduced oil production stopped. Throughout all steps of the injection procedure,
as the best method for the calculation of IFT especially for the range of the fluid flow behavior was recorded by a digital microscopic imaging
3–80 mN/m. Herein, the IFT values were calculated by a drop shape system and the obtained images were precisely analyzed by image
analysis device (DSA 100, KRUSS, Germany) which works on the base processing software to determine the variations of the phase saturation.
of pendant drop method. It has been shown that the DSA technique on
the basis of pendant drop would be able to determine the IFT values of
0.05 mN/m due to its full automatic procedure and high accuracy. The 3. Result and discussion
dynamic IFT values were also measured through connecting the IFT
apparatus to computer and running image analysis software. The results 3.1. Characterization of nanohybrids
showed that the IFT values fluctuate as a function of time till reach to a
constant value by the passage of time which is introduced as the The Raman spectroscopy is an appropriate technique for char-
equilibrium IFT in this study. acterization of carbon nanostructure specially graphene and its deri-
vatives. The Raman spectrum of the silica-graphene nanohybrid is
shown in Fig. 3. Accordingly, the D, G, and 2D bands were positioned at
2.5. Micromodel test around 1320, 1560, and 2600 cm−1, respectively. The G band is com-
monly appeared in all sp2 carbon allotropes confirming the in-plane
The fluid flow on the porous media was directly simulated and vibration of sp2 bonded carbon atoms. While the intensity of the D band
monitored in this study by using a glass micromodel. The micromodel generally shows sp3 defects which lead to increase the number of
was designed by chemical etching and laser ablation method which has crumpled graphene layers. Furthermore, the relatively high intensity of
been previously described in detail. Prior to begin each test, the glass 2D band around 2600 cm−1 that usually appears in the monolayer
micromodel was carefully washed with toluene for removing the sur- graphene sheets further verifies the creation of single layer graphene.
face contaminations and the experiments were conducted at ambient Hence, it can be concluded that the silica-graphene nanohybrid si-
temperature and pressure. After that, the pre-cleaned porous media was multaneously contains a combination of single and few layers stacked
thoroughly soaked with crude oil and the fluid was then injected at the graphene sheets.
rate of 0.03 ml/hr. The physical properties of micromodel are shown in Fig. 4 shows the TEM, HRTEM images and the related select area
Table 1. The schematic illustration of the oil-saturated micromodel and electron diffraction (SAED) patterns of silica aerogel and silica-gra-
microscopic experimental setup is presented in Fig. 2. It should be phene nanohybrid. As can be clearly seen in Fig. 4(a), size of the pri-
noted that the glass micromodel was horizontally situated for all mary silica particles is rather uniform. The high magnification TEM
flooding experiments. For oil-wet modeling, firstly, the pre-cleaned image of silica aerogel is shown in Fig. 4b; obviously, the silica aerogel
glass micromodel was shortly subjected to organic solvent such as to- mainly composed of amorphous and disordered structures. Moreover,
luene and acetone, respectively. After that, deionized water was cir- the SAED pattern of silica aerogel (Fig. 4(c)) shows only diffraction
culated through the micromodel. Subsequently, 20% HCl solution was rings without any diffraction dots confirming the silica aerogel struc-
injected and circulated to the glass micromodel for a short period of tures are amorphous. Fig. 4(d) shows the TEM image of silica-graphene
time and the deionized water was again injected to the micromodel. nanohybrid, in which a large number of wrinkled few-layered graphene
Next, the treated micromodel was dried at 200 °C for 15 min. After can be distinguished on silica aerogel. A set of highly-crumpled and
drying the micromodel, the solution of 2% trichloromethylsilane individual transparent graphene sheets confirm that these wrinkles re-
(TCMS) in toluene was injected and methylene chloride was then re- sulted from the crumpling of graphene rather than stacking, which is in
peatedly pumped and circulated into the micromodel for several times. agreement with the sharp intensity of 2D band in the Raman spectra.

255
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

Fig. 2. The schematic illustration of the oil-saturated micromodel and microscopic experimental setup.

High resolution TEM investigation of silica-graphene nanohybrids bonds of the functionalized samples were noticeably intensified com-
shows that the thin walls typically consist of only few layers of gra- pare to pristine silica-graphene nanohybrid. Furthermore, by increasing
phene sheets formed on silica aerogel (Fig. 4(e)). As further evidence, the time of functionalization the intensity of O1s bands were com-
the crystalline structure of graphene on silica was verified by selected paratively raised indicating the creation of more oxygen functional
area electron diffraction (SAED), as shown in Fig. 4(f). A ring-like dif- groups on silica-graphene nanohybrids (Fig. 5). The C1s spectra of si-
fraction pattern verified the loss of ordering in the sheets [33]. lica-graphene nanohybrids were fitted with four Gaussian peaks related
X-ray photoelectron spectrum (XPS) was employed to determine the to CeC, CeO, C]O, and O]CeO binds (see figure S2). The values of
element composition and oxidation state of the prepared silica-gra- carbon to oxygen (C/O) ratio of the samples were calculated and
phene and functionalized silica-graphene nanohybrids. The wide survey summarized in Table S1. The XPS spectra demonstrate noticeable dif-
of XPS spectra reveal that all of the pristine and functionalized silica- ferences between pristine and functionalized silica-graphene nanohy-
graphene nanohybrids only composed of carbon, oxygen, and silica as brids as the proportion of oxygen to carbon augments by increasing the
the main elements indicating the lack of any impurity. Fig. 5 illustrates time of functionalization from 24 to 72 h. However, by changing
that all of the spectra contain four distinctive peaks in the vicinity of functionalization time there is no distinct alteration regarding the type
∼101, 151, 283, and 531 eV which are related to Si 2p, Si 2 s, C1s, and of oxygen containing functional groups (see Table S1).
O1s bonds, respectively. As observed, the intensity of the Si 2p, Si 2 s,
and C1s bonds remain unchanged in all samples. However, the O1s
3.2. Phase behavior of functionalized nanohybrids dispersions in presence
of NaCl

The visual observations of stability of fsg.72 nanohybrid, as an ex-


ample, are shown in Figs. 6 and 7. The addition of electrolytes in
aqueous phase increases the thickness of the electrical double layer and
causes the destabilization of nanoparticles dispersions. Fig. 6 shows the
phase behavior of 0.1 wt% fsg.72 dispersions that are various in NaCl
concentrations. As can be seen in Fig. 6, the addition of NaCl up to 4 wt
% does not lead to destabilize the aqueous dispersion of nanohybrids.
This observation suggests that 4 wt% NaCl represents a critical salt
concentration above which nanohybrid aggregation and sedimentation
visually occurs. Moreover, Fig. 7 shows the phase behavior of fsg.72
dispersions in different NaCl and nanoparticle concentrations. These
observations show that the precipitation and sedimentation of nano-
particles occur by either increasing NaCl or nanoparticles concentra-
tion. It is obvious from Fig. 7(a–c) that the fsg.72 nanohybrid disper-
sions with 0.01, 0.05, and 0.1 wt% nanoparticle concentrations are
stable for NaCl concentration up to 4 wt%. However, for 0.15 wt%
fsg.72 nanohybrid, the small addition of NaCl even 1 wt% leads to
destabilize the aqueous dispersion and settle of nanoparticles (see
Fig. 7(d)). Therefore, regarding to these observations, the nanohybrids
concentrations up to 0.1 wt% and 4 wt% NaCl concentration were se-
Fig. 3. Raman spectrum of silica-graphene nanohybrid. lected for more investigation in this study.

256
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

Fig. 4. TEM images of (a) silica aerogel, and (d) silica-graphene nanohybrids, HRTEM images of (b) silica aerogel, and (e) silica-graphene nanohybrids and the
corresponding SAED patterns of (c) silica aerogel, and (f) silica-graphene nanohybrids.

Fig. 5. X-ray photoelectron spectra of (a) silica-graphene nanohybrid, (b)


fsg.24, (c) fsg.48, and (d) fsg.72.

3.3. Mechanism investigation of functionalized silica-graphene nanofluid in


EOR

Generally, one of the prominent mechanisms has been proposed for Fig. 6. Phase behavior of 0.1 wt% fsg.72 nanohybrid dispersions at various
NaCl concentrations at different time span (a) 1 h and (b) 24 h after dispersion.
enhanced oil recovery when nanoparticles add to the injected fluid is
reducing the interfacial tension between the nanofluid and oil [34]. In
fact, the adsorption of nanoparticles may occur at the oil-water inter- 3.3.1. IFT measurements
face conducing the reduction of IFT [35]. The IFT values of functionalized silica-graphene nanohybrids at
different concentrations (0.01, 0.05, and 0.1 wt %) were measured

257
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

Fig. 7. Phase behavior of various concentrations of fsg.72 nanohybrid dispersions (a) 0.01 wt%, (b) 0.05 wt%. (c) 0.1 wt% and (d) 0.15 wt% at various NaCl
concentrations (1 h after dispersion).

Fig. 8. The effect of the concentration functionalized silica-graphene nanohybrids on the IFT values.

through DSA mechanism as previously described. Fig. 8 shows the drop phase, respectively. The IFT value between the crude oil and
variation of IFT values with increasing the concentration of the func- deionized water was calculated to be about 19.11 mN/m. As can be
tionalized nanohybrids at various functionalization time. In the ex- seen in Fig. 8, at the concentration of 0.1 wt%, the IFT value reaches to
periments, the dispersants and crude oil were applied as the bulk and ∼10.44, 9.01, and 7.13 mN/m for fsg.24, fsg.48, and fsg.72 nanofluids,

258
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

Fig. 11. The effect of different functionalized silica-graphene nanohybrids on


Fig. 9. Comparison of the IFT values of functionalized silica-graphene nano- the average diameter of emulsions droplets at the concentration of 0.1 wt%
hybrids in deionized water and synthetic brine (4 wt% NaCl) at the con- functionalized nanohybrids in deionized water and brine solution (24 h after
centration of 0.1 wt%. preparation of emulsions).

respectively; suggesting that the IFT value can be reduced more than fsg.48 which can be attributed to the better dispersion and higher sta-
60% in the presence of functionalized silica-graphene nanohybrids. It bility of this sample versus fsg.24 and fsg.48 due to its higher con-
can be clearly observed that the IFT decreases with increasing the centration of functional groups. Indeed, functional groups on the edge
concentration of the functionalized nanohybrids which can be related plane and surface of the nanohybrids can impede the particle-particle
to the adsorption of functionalized nanohybrids on the oil-water in- interaction through decreasing the aggregation of nanoparticles;
terface. As a matter of fact, the functionalized nanohybrids create an therefore, higher amount of nanohybrids can migrate to the oil-water
extra layer at the interface of the oil and nanofluid which extends the interface which reduces IFT [37]. Fig. 9 exhibits the ionic strength of
interface through acting as an amphiphilic surfactant. Hence, the ad- nanofluids at the concentration of 0.1 wt% functionalized silica-gra-
dition of functionalized nanohybrids noticeably reduces the capillary phene nanohybrids by comparing the IFT values in deionized water and
forces and consequently increases the capillary number [36]. Further- brine solution (4 wt % NaCl). It should be mentioned that the presence
more, in the case of fsg.72, the IFT values reduced more than fsg.24 and of salt may alter the oil-water interface through changing the

Fig. 10. The microscopy images of the emulsions at the concentration of 0.1 wt% functionalized nanohybrids; (a) fsg.24, (b) fsg.48, and (c) fsg.72 in deionized water,
(d) fsg.24, (e) fsg.48, and (f) fsg.72 in brine solution. (The images were taken 24 h after preparation of emulsions).

259
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

nanoparticles, respectively. The partition coefficients of the functiona-


lized nanohybrids between decalin and water were experimentally
determined through dispersion of functionalized nanohybrids (0.1 wt
%) in equal volume of deionized water (Fig. 10(a–c)) and brine solu-
tion-decalin emulsions (Fig. 10(d–f)). In the case of deionized water, the
lower phase shows light-brown solution due to the partially dispersion
of functionalized silica-graphene nanohybrids (left side of Fig. 10(a–c)).
Contrary, by adding salt, an absolute thin colorless phase appears after
24 h indicating that the sodium cations attracted the functionalized
nanohybrids toward the interface by electrostatic interactions (right
side of Fig. 10(a–c)). Thermodynamically, the irreversible adsorption of
nanohybrids to the oil-water interface creates more stable emulsion.
furthermore, the high stability of emulsions by the functionalized silica-
graphene nanohybrids can be ascribed to electrostatic repulsions and
steric hindrance through the formation of some rigid shells which oc-
curs by the adsorption of nanoparticles on the interface [44,45]. Ob-
viously, spherical droplets with various diameters can be seen in
deionized water and brine emulsions. The average diameter of emulsion
droplets in deionized water was approximately estimated to be about
0.45–1.06 μm which is extremely smaller than silica-carbon nanotube
nanohybrids and functionalized graphene [27,46]. As presented in
Fig. 10, the microscopic images reveal that in deionized water, most of
Fig. 12. The behavior of an added droplet of emulsion in deionized water and the droplets possess smaller diameter with better distribution than
decalin. those of brine emulsions due to less electrostatic repulsion in the pre-
sence of salt. According to the possible interactions in the proposed
hydrophilic–lipophilic balance and compressing the electrical double emulsions, the hydrophobic parts of nanohybrids have more tendencies
layer at the same time. The consequence of these phenomena can re- to interact with oil; however, the oxygen containing functional groups
duce the solidity of the interface trough increasing the concentration of remain in aqueous phase due to their hydrophilic properties. In spite of
nanoparticles at the oil-water interface which leads to the reduction of the closeness of the emulsion droplets in microscopic images, uniform
IFT [38]. From other point of view, by adding NaCl, there is an op- distribution of emulsion droplets further emphasizes the role of func-
portunity for positive and negative ions to form a closer contact with tional groups as a hydrophilic surfactant which improves the stability.
oil-water interface. As it happens, the sodium cations can easily interact The effect of functionalization time on the average diameter of
with the negative charged functional groups on the surface of functio- emulsion droplets was also investigated. As can be seen in Fig. 11, the
nalized silica-graphene nanohybrids. Thus, the negative charged func- average diameter of emulsion drops decreases by increasing the func-
tional groups will be surrounded by Na+ which reduces the double tionalization time indicating better dispersion of nanohybrids, uniform
layer thickness and increases the density of nanoparticles at the oil- distribution of droplets, and higher stability of emulsions by increasing
water interface leading to the reduction of IFT. Considering the pre- the amount of oxygen containing functional groups. In addition, to
sence of anions in the double layer, they can be also adsorbed to the oil- observe individual behavior of emulsion in deionized water and dec-
water interface by electrostatic interactions. However, the adsorption of alin, an equal amount of emulsion was dropped in a constant volume of
chloride anions is assumed to occur at the hydrophobic-hydrophilic deionized water and decalin. Generally, when a drop of oil-in-water
interfaces which have been confirmed by molecular dynamics simula- emulsion disperses in water, the added drop remains unchanged in oil.
tion [39]. Accordingly, as can be seen in Fig. 9, the IFT values decrease By adding a drop of the as-prepared emulsions to water and decalin, it
by increasing the ionic strength which highlights the augmentation of was found that the type of emulsions was decalin in water (see Fig. 12
nanoparticles at the oil-water interface. Therefore, the addition of as an example).
functionalized silica-graphene nanohybrids to the brine solution might Nanoparticles adsorption at the oil and water interface, regard to
facilitate the EOR process. Also, the results indicate that the minimum their high surface area, and their emulsions, can cause efficacious
value of IFT can be reached at the highest concentration of nanohybrids treatment on heavy oil components, due to their polar agents. As an
(0.1 wt%). Hence, it was selected as the optimum concentration of issue incumbent to be seen, the effect of the nanoparticles on creating
nanofluid for further investigation of the average droplets diameter in Pickering emulsion and its behavior during the increase of the con-
Pickering emulsion, and micromodel. centration of the asphaltene, have been investigated. Figs. 13, and 14
display the microscopy images of emulsions which were prepared by
0.1 wt% functionalized nanohybrids in brine and different concentra-
3.3.2. Emulsion tions of asphaltene in toluene with their corresponding droplet size
To reach a stable Pickering emulsion, the functionalized silica-gra- distribution plots, respectively. The long chain of the asphaltene mo-
phene nanohybrids must adsorb at the oil-water interface and occupy lecules caused nanoparticles to migrate to the outer shell of the as-
an enormous interfacial area to hinder droplets coalescence. Physically, phaltene molecules and share their hydrophibic chain, while nano-
the creation of oil-water interface in the presence of nanoparticles particles hydrophilic head still remained in water, and create large scale
needs an extra force to subdue the Laplace pressure which is provided emulsions. On the other hand, the small size of emulsions, due to the
by ultrasonic waves [40]. Generally, the stability of droplets in Pick- Brownian motion of the particles, redistributes the particles to an
ering emulsions firmly depends on the level of adsorption energy of equilibrium configuration with the lowest total free energy, and
nanoparticles and their surface coverage at the interface [41–43]. The minimum possible IFT. As can be seen in Fig. 14(a–c), a shift towards a
functionalized silica-graphene nanohybrids enhance the stability of broader droplet size distribution was observed with increase in as-
emulsions by taking the advantage of hydrophilic and hydrophobic phaltene concentration in all emulsions. Obviously, the average dro-
characteristics emanating from functional groups and silica-graphene plets diameter of emulsions for all samples, due to the heavy organic

260
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

Fig. 13. The microscopy images of the emulsions at the 0.1 wt% of (a) fsg.24, (b) fsg.48, (c) fsg.72 with different concentrations of asphaltene (1). 0.01 wt%, (2)
0.1 wt%, (3) 0.5 wt%, (4) 1 wt% (The images were taken 24 h after preparation of emulsions).

261
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

Fig. 14. Histograms of the droplet size distribution as a function of asphaltene concentration: (a) fsg.24, (b) fsg.48, (c) fsg.72, and (d) average droplets diameter as a
function of asphaltene concentration (concentration of nanohybrids = 0.1 wt%).

molecules, long-chain components and their polar character, was in- shows the differential pressure and oil recovery of the micromodels.
creased with increase in asphaltene concentration. Furthermore, the The final stage of water flooding in which the oil production is being
concentration of the brine in the water, because of ions existed in the stopped is presented in Fig. 15(a). As it can be seen in Fig. 16(a), during
water and stand along the nanoparticles and then emulsions would be water injection, the pressure difference between injecting and produ-
the second reason of the diameter enhancement of the emulsions. The cing ports, is growing up until the first droplets of injected water in
stability of the emulsion would decrease with increase in emulsion micromodel was produced. The figure also shows that the growing in-
droplets size. crease of the oil production was occurred before that point, and after
Asphaltene precipitation and the adsorption of nanoparticles in that the oil production decreased gradually until it stopped. It was also
porous media can reduce its porosity and permeability. Additionally, obvious that in water injection, the water flow has more tendencies for
emulsion trapping thorough pore throats can reduce rock permeability. passing through the most permeable and shortest path due to the fin-
Therefore, the emulsion droplets magnitude could be an important gering phenomenon. Therefore, water flowed in the areas with the
issue which may damage porous media, and should be investigated in minimum resistance in which no or low oil recovery has been achieved
presence of the heavy oil components to find out the worst possible and consequently reduced the EOR performance. Regarding the IFT
effects on the emulsion droplets diameter. As a conclusion, the con- values between oil and water, the highest IFT values leads to the lowest
centration of the asphaltene has a negative effect on the emulsion sta- capillary number [53]. The results show that only 31% of oil can be
bilization and droplets. However, nanoparticles investigated in this recovered by water flooding which is relatively poor. It means that
study has not raised droplets diameter so much, and could be called that water partially swept oil as the oil saturation reaches to 69%, hence, a
no damage would occur during the injection because of the droplets size large amount of oil was remained at the end of displacement. Conse-
of the emulsions. quently, the recovery of the oil is increased until the reducing pressure
difference in the micromodel became stable and there was no change in
4. Micromodel tests pushing force to produce more oil.
The recovery of water flooding and after that, nanofluids flooding
The micromodels of water flooding and nanofluid flooding at the are conducted at the concentration of 0.1 wt% functionalized nanohy-
concentration of 0.1 wt% functionalized nanohybrids in brine in the last brids in brine. Water and nanofluids flooding experiments were carried
moment of the test are shown in Fig. 15(a–d). Moreover, Fig. 16(a–d) out with the flow rate of 0.0005cc/min. The final stage of water

262
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

Fig. 15. Micromodel after (a) water flooding, and nanofluid flooding of (b) fsg.24, (c) fsg.48, and (d) fsg.72. (All nanofluids were prepared in synthetic brine solution
at the concentration of 0.1 wt% functionalized nanohybrids).

flooding in a separate test, which the oil production is being stopped, is viscosity enhancement on the nanofluid injected in the micromodel.
presented in Fig. 15(a). All nanofluid tests were taken after water Consequently, according to our calculations and experimental ob-
flooding and secondary recovery, to investigate tertiary recovery. The servations, fsg.72 as the best functionalized nanohybrid exhibits better
micromodels (Fig. 15(b–d)) illustrate the oil production responses EOR performances compared to fsg.48, and fsg.24, respectively.
during the water flooding and nanofluids flooding of fsg.24, fsg48, and
fsg72.
Comparatively, the enhancement of macroscopic sweep efficiency in 5. Conclusion
the presence of functionalized nanohybrids can be observed by in-
creasing the amount of oxygen containing functional groups. The oil In summary, the silica-graphene nanohybrid was prepared via CVD
recovery from nanofluid flooding was calculated to be about 53, 71, method and functionalized by using a mixture of nitric and sulfuric acid
and 83% for fsg.24, fsg.4, and fsg.72, respectively (see Fig. 16(b–d)), vapors. The functionalization process was carried out in different time
further confirming the effect of oxygen containing functional groups on periods. The XPS spectra show that the augmentation of defect struc-
the EOR performance. This increase in EOR performance which oc- tures and the oxygen containing functional groups is due to the passage
curred in high functionalized nanofluids must be due to those changes of functionalization time [48]. Then, the potential application of
which affect the injecting fluid properties and enhanced them. Ac- functionalized silica-graphene nanohybrids for enhanced oil recovery
cording to the IFT figures, in pervious section it can be concluded that was perfectly investigated. At first step, the effect of ionic strength and
the amount of oxygen containing functional groups affects the inter- the concentration of functionalized silica-graphene on interfacial ten-
facial tension between oil and water. The low interfacial tension in sion between crude oil and water were studied. The results reveal that
highly functioned nanoparticles is due to the hydrophilicity effects of IFT values can be reduced more than 60% in the presence of functio-
the highly functioned nanoparticles groups, and make them amphi- nalized silica-graphene nanohybrids. The stabilization of oil in water
philic particles; solute in water and oil. emulsions was also investigated at the concentration of 0.1 wt% func-
As a result, it can be concluded that, the higher the functional tionalized silica-graphene nanohybrids in deionized water and 4 wt%
groups embedded, the lower the interfacial tension resulted, and ob- NaCl brine solution. The results exhibit that the electrostatic interac-
viously, the higher the recovery enhanced in micromodel. Apparently, tions between positively charged sodium and negatively charged
the reduction of IFT influences the contact between by passed oil and oxygen containing functional groups may lead to better stability of the
injected fluid. When the nanofluids were injected after waterflooding; emulsions. Finally, the micromodel visualizations demonstrate that oil
the reduction of IFT values leads to better mobilization and more oil recovery noticeably enhances when nanofluid flooding occurs after
production. This occurred as a result of the increase in oxygen con- water flooding. It goes without saying that employing functionalized
taining functional groups, embedded on the nanoparticles structural silica-graphene nanohybrids even at low concentration in this study not
shell. As it can be seen in Fig. 16(a–d), gradually decrease in pressure only improves oil recovery parameters, but also introduces an economic
difference in micromodels, means that there is no or negligible effect of alternative for nanoadditive in reservoirs.

263
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

Fig. 16. Oil recovery performance, differential pressure vs. injected pore volume of micromodel after (a) water flooding, and nanofluid flooding of (b) fsg.24, (c)
fsg.48, and (d) fsg.72.

Acknowledgement Nanotechnology Conference and Exhibition (2012).


[10] O. Torsaeter, S. Li, L. Hendraningrat, Enhancing oil recovery of Low-permeability
berea sandstone through optimised nanofluids concentration, SPE Enhanced Oil
This research was supported by Iran National Science Foundation Recovery Conference (2013).
for postdoctoral fellowship with number of 96006744. [11] L. Hendraningrat, O. Torsæter, Metal oxide-based nanoparticles: revealing their
potential to enhance oil recovery in different wettability systems, Appl. Nanosci. 5
(2) (2015) 181–199.
Appendix A. Supplementary data [12] B. Ju, T. Fan, M. Ma, Enhanced oil recovery by flooding with hydrophilic nano-
particles, China Particuol. 4 (01) (2006) 41–46.
Supplementary material related to this article can be found, in the [13] C. Drummond, J. Israelachvili, Surface forces and wettability, J. Pet. Sci. Eng. 33 (1)
(2002) 123–133.
online version, at doi:https://doi.org/10.1016/j.colsurfa.2018.08.029. [14] C.C. Agbalaka, et al., The effect of wettability on oil recovery: a review, SPE Asia
Pacific Oil and Gas Conference and Exhibition (2008).
References [15] K. Xu, et al., Microfluidic investigation of nanoparticles’ role in mobilizing trapped
oil droplets in porous media, Langmuir 31 (51) (2015) 13673–13679.
[16] P. Stevenson, Dimensional analysis of foam drainage, Chem. Eng. Sci. 61 (14)
[1] T. Austad, Water-based EOR in carbonates and sandstones: new chemical under- (2006) 4503–4510.
standing of the EOR potential using smart water, Enhanced Oil Recov. Field Case [17] N.Y.T. Le, et al., Design and screening of synergistic blends of SiO2 nanoparticles
Stud. (2013) 301–335. and surfactants for enhanced oil recovery in high-temperature reservoirs, Adv. Nat.
[2] J. Zuta, I. Fjelde, Mechanistic modeling of CO2-foam processes in fractured chalk Sci. Nanosci. Nanotechnol. 2 (3) (2011) 035013.
rock: effect of foam strength and gravity forces on oil recovery, SPE Enhanced Oil [18] O. Torsater, S. Li, L. Hendraningrat, Effect of some parameters influencing en-
Recovery Conference (2011). hanced oil recovery process using silica nanoparticles: an experimental investiga-
[3] D.W. Green, G.P. Willhite, Enhanced Oil Recovery, Richardson, Tex.: Henry L. tion, SPE Reservoir Characterization and Simulation Conference and Exhibition
Doherty Memorial Fund of AIME, Society of Petroleum Engineers, 1998. (2013).
[4] H. Pei, et al., Comparative effectiveness of alkaline flooding and alkaline–surfactant [19] A. Maghzi, et al., Pore-scale monitoring of wettability alteration by silica nano-
flooding for improved heavy-oil recovery, Energy Fuels 26 (5) (2012) 2911–2919. particles during polymer flooding to heavy oil in a five-spot glass micromodel,
[5] M. Dong, S. Ma, Q. Liu, Enhanced heavy oil recovery through interfacial instability: Transp. Porous Media 87 (3) (2011) 653–664.
a study of chemical flooding for Brintnell heavy oil, Fuel 88 (6) (2009) 1049–1056. [20] N. Ogolo, O. Olafuyi, M. Onyekonwu, Enhanced oil recovery using nanoparticles,
[6] N. Lai, et al., A water‐soluble acrylamide hydrophobically associating polymer: SPE Saudi Arabia Section Technical Symposium and Exhibition (2012).
synthesis, characterization, and properties as EOR chemical, J. Appl. Polym. Sci. [21] J. Shi, et al., Synergy of pickering emulsion and sol‐gel process for the construction
129 (4) (2013) 1888–1896. of an efficient, recyclable enzyme cascade system, Adv. Funct. Mater. 23 (11)
[7] Nielson, J.P., Recovery of oil from oil-bearing formation by continually flowing (2013) 1450–1458.
pressurized heated gas through channel alongside matrix. 1989, Google Patents. [22] T. Chen, P.J. Colver, S.A. Bon, Organic–Inorganic Hybrid Hollow Spheres Prepared
[8] G.A. Pope, The application of fractional flow theory to enhanced oil recovery, Soc. from TiO2‐Stabilized Pickering Emulsion Polymerization, Adv. Mater. 19 (17)
Pet. Eng. J. 20 (03) (1980) 191–205. (2007) 2286–2289.
[9] C.R. Miranda, L.Sd. Lara, B.C. Tonetto, Stability and mobility of functionalized si- [23] C. Wu, et al., Nanoparticle cages for enzyme catalysis in organic media, Adv. Mater.
lica nanoparticles for enhanced oil recovery applications, SPE International Oilfield

264
S. Tajik et al. Colloids and Surfaces A 556 (2018) 253–265

23 (47) (2011) 5694–5699. [35] B.P. Binks, Particles as surfactants—similarities and differences, Curr. Opin. Colloid
[24] J. Zhou, et al., Magnetic Pickering emulsions stabilized by Fe3O4 nanoparticles, Interface Sci. 7 (1) (2002) 21–41.
Langmuir 27 (7) (2011) 3308–3316. [36] M.J. Rosen, et al., Ultralow interfacial tension for enhanced oil recovery at very low
[25] X.-C. Luu, J. Yu, A. Striolo, Nanoparticles adsorbed at the water/oil interface: surfactant concentrations, Langmuir 21 (9) (2005) 3749–3756.
coverage and composition effects on structure and diffusion, Langmuir 29 (24) [37] M.S. Manga, et al., Measurements of submicron particle adsorption and particle film
(2013) 7221–7228. elasticity at oil–Water interfaces, Langmuir 32 (17) (2016) 4125–4133.
[26] S. Crossley, et al., Solid nanoparticles that catalyze biofuel upgrade reactions at the [38] L. Han, et al., The interfacial tension between cationic gemini surfactant solution
water/oil interface, Science 327 (5961) (2010) 68–72. and crude oil, J. Surfactants Deterg. 12 (3) (2009) 185–190.
[27] M. Shen, D.E. Resasco, Emulsions stabilized by carbon nanotube− silica nanohy- [39] C. Tian, Y. Shen, Structure and charging of hydrophobic material/water interfaces
brids, Langmuir 25 (18) (2009) 10843–10851. studied by phase-sensitive sum-frequency vibrational spectroscopy, Proc. Natl.
[28] M. Buchgraber, et al., Creation of a dual-porosity micromodel for pore-level vi- Acad. Sci. 106 (36) (2009) 15148–15153.
sualization of multiphase flow, J. Pet. Sci. Eng. 86 (2012) 27–38. [40] K.Y. Yoon, et al., Graphene oxide nanoplatelet dispersions in concentrated NaCl and
[29] M. Buchgraber, A.R. Kovscek, L.M. Castanier, A study of microscale gas trapping stabilization of oil/water emulsions, J. Colloid Interface Sci. 403 (2013) 1–6.
using etched silicon micromodels, Transp. Porous Media 95 (3) (2012) 647–668. [41] T.S. Horozov, B.P. Binks, Particle‐stabilized emulsions: a bilayer or a bridging
[30] S. Tajik, B. Nasernejad, A. Rashidi, Preparation of silica-graphene nanohybrid as a monolayer? Angew. Chemie Int. Ed. 45 (5) (2006) 773–776.
stabilizer of emulsions, J. Mol. Liq. 222 (2016) 788–795. [42] F. Kim, L.J. Cote, J. Huang, Graphene oxide: surface activity and two‐dimensional
[31] U.K. Bangi, S.L. Dhere, A.V. Rao, Influence of various processing parameters on assembly, Adv. Mater. 22 (17) (2010) 1954–1958.
water-glass-based atmospheric pressure dried aerogels for liquid marble purpose, J. [43] J. Kim, et al., Graphene oxide sheets at interfaces, J. Am. Chem. Soc. 132 (23)
Mater. Sci. 45 (11) (2010) 2944–2951. (2010) 8180–8186.
[32] S. Tajik, B. Nasernejad, A. Rashidi, Surface modification of silica-graphene nano- [44] P. Guo, H. Song, X. Chen, Hollow graphene oxide spheres self-assembled by W/O
hybrid as a novel stabilizer for oil-water emulsion, Korean J. Chem. Eng. 34 (9) emulsion, J. Mater. Chem. 20 (23) (2010) 4867–4874.
(2017) 2488–2497. [45] H.A. Wege, et al., Long-term stabilization of foams and emulsions with in-situ
[33] A. Amiri, et al., Microwave-assisted synthesis of highly-crumpled, few-layered formed microparticles from hydrophobic cellulose, Langmuir 24 (17) (2008)
graphene and nitrogen-doped graphene for use as high-performance electrodes in 9245–9253.
capacitive deionization, Sci. Rep. 5 (2015). [46] Y. He, et al., Factors that affect pickering emulsions stabilized by graphene oxide,
[34] G. Cheraghian, L. Hendraningrat, A review on applications of nanotechnology in the ACS Appl. Mater. Interfaces 5 (11) (2013) 4843–4855.
enhanced oil recovery part B: effects of nanoparticles on flooding, Int. Nano Lett. 6 [48] M. Hashemi, et al., Functionalized R9–reduced graphene oxide as an efficient nano-
(1) (2016) 1–10. carrier for hydrophobic drug delivery, RSC Adv. 6 (78) (2016) 74072–74084.

265

You might also like