You are on page 1of 6

Nanofluid of graphene-based amphiphilic Janus

nanosheets for tertiary or enhanced oil recovery:


High performance at low concentration
Dan Luoa,b,c,1, Feng Wanga,1, Jingyi Zhuc, Feng Caoa, Yuan Liua, Xiaogang Lic, Richard C. Willsonb,d, Zhaozhong Yangc,2,
Ching-Wu Chua,e,2, and Zhifeng Rena,2
a
Department of Physics and Texas Center for Superconductivity, University of Houston, Houston, TX 77204; bDepartment of Chemical and Biomolecular
Engineering, University of Houston, Houston, TX 77204; cState Key Laboratory of Oil and Gas Reservoir Geology and Exploitation, Southwest Petroleum
University, Chengdu, Sichuan 610500, China; dDepartamento de Biotecnología e Ingeniería de Alimentos, Tecnologico de Monterrey, Monterrey,
Nuevo Leon 64849, Mexico; and eLawrence Berkeley National Laboratory, Berkeley, CA 94720

Contributed by Ching-Wu Chu, May 25, 2016 (sent for review April 28, 2016; reviewed by Caili Dai and Dapeng Yu)

The current simple nanofluid flooding method for tertiary or flooding method for tertiary oil recovery that is comparable to the
enhanced oil recovery is inefficient, especially when used with sophisticated chemical methods. We anticipate that this work will
low nanoparticle concentration. We have designed and produced bring simple nanofluid flooding at low concentration to the stage
a nanofluid of graphene-based amphiphilic nanosheets that is very of oil field practice, which could result in oil being recovered in a
effective at low concentration. Our nanosheets spontaneously more environmentally friendly and cost-effective manner.
approached the oil–water interface and reduced the interfacial
tension in a saline environment (4 wt % NaCl and 1 wt % CaCl2), Materials and Methods

APPLIED PHYSICAL
regardless of the solid surface wettability. A climbing film Synthesis. The Janus amphiphilic nanosheets were produced by tuning the Janus

SCIENCES
appeared and grew at moderate hydrodynamic condition to en- balance of graphene oxide (GO) with alkylamine. Initially, GO was synthesized
capsulate the oil phase. With strong hydrodynamic power input, a from chemical oxidation of graphite (15). Single-surface hydrophobization was
solid-like interfacial film formed and was able to return to its orig- then carried out (16, 17). The nanofluid was made stable to avoid agglomer-
inal form even after being seriously disturbed. The film rapidly ation of the nanosheets. Brine used in all experiments contained 4 wt % NaCl
separated oil and water phases for slug-like oil displacement. and 1 wt % CaCl2.
The unique behavior of our nanosheet nanofluid tripled the best
Characterization. Atomic force microscopy (AFM; Veeco Dimensions 3000 atomic
performance of conventional nanofluid flooding methods under
force microscope) was used to examine the morphology of GO. Measurement was
similar conditions.
conducted using silicon AFM probes (HQ:NSC15/AL BS, Mikromasch) with a resonant
frequency of ∼325 kHz, a force constant of ∼40 N m−1, and a tip radius of ∼8 nm.
|
nanofluid flooding amphiphilic Janus nanosheets | enhanced oil Imaging was done in tapping mode with resolutions of 512 × 512. Scanning electron
| |
recovery climbing film interfacial film microscopy (SEM; FEI Quanta 200) was used to examine the cross-section of the
sandstone cores under an accelerating voltage of 20 kV. Fourier transform infrared

F inding economically viable and environmentally friendly


methods to extract the huge amount of residual oil after
primary and secondary recovery remains challenging for the oil
spectroscopy (FTIR) spectra were recorded on a Nicolet iS50 FTIR spectrometer with
an attenuated total reflectance accessory. Thermal gravimetric analysis (TGA) was

and gas industry and is also of significant importance in efforts to Significance


satisfy the world’s increasing energy demand. Nanofluid flooding
as an alternative tertiary oil recovery method has been recently Improving crude oil recovery by 1% worldwide would result in
reported (1–5). Obviously, simple nanofluid flooding (containing a huge amount of crude oil resources becoming available.
only nanoparticles) at low concentration (0.01 wt % or less) However, the economic and environmental concerns are too
shows the greatest potential from the environmental and eco- serious to ignore when chemical methods (surfactants or
nomic perspective. Several corresponding oil displacement polymers flooding, etc.) are used for an average 10–20% en-
mechanisms have also been introduced, including reduction of hancement for tertiary oil recovery. Simple nanofluid (con-
oil–water interfacial tension (6, 7), alteration of rock surface taining only nanoparticles) flooding at low concentration
wettability (8–10), and generation of structural disjoining pres- (0.01 wt % or less) is a promising alternative, but the efficiency
sure (11–13). However, the oil recovery factor is below 5% with is below 5% in a saline environment (2 wt % or higher NaCl
0.01% nanoparticle loading in core flooding tests in a saline content). We report a simple nanofluid of graphene-based Ja-
environment (2 wt % or higher NaCl content). Here we show nus amphiphilic nanosheets for enhanced oil recovery with
that an oil recovery factor of 15.2% is achieved by using a simple efficiency of about 15.2%, comparable to chemical methods,
nanofluid of graphene-based Janus amphiphilic nanosheets. which is both economically and environmentally beneficial to the
To our knowledge, this is the first report of applying nanofluid of petroleum industry.
amphiphilic Janus two-dimensional materials in tertiary or en-
hanced oil recovery. We found that in a saline environment, the Author contributions: D.L. and Z.R. conceived the original idea; D.L., F.W., R.C.W., and Z.Y.
designed research; D.L., F.W., J.Z., F.C., Y.L., and X.L. performed research; D.L., F.W., F.C.,
nanosheets spontaneously approach the oil–water interface, re- Y.L., and R.C.W. analyzed data; Z.Y., C.-W.C., and Z.R. supervised the project; and D.L.,
ducing the interfacial tension. A climbing film emerges and en- F.W., C.-W.C., and Z.R. wrote the paper.
capsulates the oil phase and may carry it forward. Furthermore, Reviewers: C.D., China University of Petroleum, Huadong Campus; and D.Y., Peking
we found that a solid-like film forms with strong hydrodynamic University.
power. The film rapidly separates oil and water for slug-like oil The authors declare no conflict of interest.
displacement. Even though there are ways to achieve 20% en- 1
D.L. and F.W. contributed equally to this work.
hanced recovery by complicated alkali/surfactant/polymer flood- 2
To whom correspondence may be addressed. Email: Yangzhaozhong@swpu.edu.cn, cwchu@
ing (14) or by surfactants with added nanoparticles (5), the uh.edu, or zren@uh.edu.
necessary concentrations of the chemicals and nanoparticles This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
are much higher than 0.01 wt %. Our results provide a nanofluid 1073/pnas.1608135113/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1608135113 PNAS | July 12, 2016 | vol. 113 | no. 28 | 7711–7716


effect (Fig. 1). After settling for 2 h, unmodified GO precipitated
while the interfacial film of our nanosheets remained intact. This
observation suggested successful asymmetrical functionalization of
GO with hydrocarbon chains. In FTIR analysis (Fig. 2A), both
GO and amphiphilic nanosheets exhibited peaks at 1,723 cm−1,
1,587 cm−1, and 1,230 cm−1, which can be assigned to C=O car-
bonyl/carboxyl, C=C aromatic, and C–O–C epoxy vibrations,
respectively. In addition, amphiphilic nanosheets clearly dem-
onstrated the presence of methylene groups (strong peaks at
1,470 cm−1, 2,850 cm−1, and 2,925 cm−1) and methyl groups
Fig. 1. Behaviors of unmodified GO and amphiphilic nanosheets in the (weaker signals at 1,380 cm−1 and 2,960 cm−1), which confirmed
heptane/brine system. Small pieces of interfacial film were attached to the the successful conjugation of hydrocarbon chains onto the GO
hydrophilic glass surface in the heptane phase due to its amphiphilicity, surface. Compared with GO, the TGA curve of amphiphilic
which appeared as black dots. nanosheets displayed an additional weight loss stage between
400 °C and 500 °C, which may be attributed to the decomposition
performed on a TGA Q50 (TA Instrument) under nitrogen atmosphere at a rate of
of the carbon chains (Fig. 2B).
10 °C/min. Particle size and concentration were detected and visualized using a
Malvern NanoSight NS300. Stability Evaluation of the Nanofluid. It is crucial that the nanofluid
has good stability before being injected into the reservoir. From
Core Flooding Test. Four man-made sandstone rock cores were tested in the observation, both GO and amphiphilic nanosheets have very small
flooding equipment (Fig. S1). The physical properties of the rock cores were amounts of precipitates even after 30 d. To evaluate the stability in
measured and are listed in Table S1. Crude oil samples were taken from one of microscopic view, the dispersions of GO and amphiphilic nano-
China’s oil fields. The viscosity of the crude oil was 75 cP at 25 °C. Nanofluids of sheets were first subjected to bath sonication for 30 s before dilution
0.005 wt % and 0.01 wt % nanosheet concentration were injected into a saline to the concentration of 0.005 wt % and then injected into the
environment (4 wt % NaCl and 1 wt % CaCl2) to measure the enhanced oil chamber in the Malvern NS300 system, in which a laser passes
recovery factor for rock cores with both low (samples 1 and 2) and high (samples through. Particles in the path of the laser scattered light and were
3 and 4) liquid permeability. The core flooding test was conducted sequentially visualized by a camera. Altogether, five locations were observed,
with the following steps: cleaning of rock cores; saturating cores with brine;
each for a period of 60 s. Fig. 3 shows the average hydrodynamic
establishing initial brine water and oil saturation by oil injection until no more
diameter distribution of GO (Fig. 3A) and amphiphilic nanosheets
brine water was produced; brine water flooding until no more oil (i.e., 100%
water cut) was produced; and nanofluid flooding until no more oil was
(Fig. 3B). It appeared that the mode of hydrodynamic diameter of
extracted. The total injection volume of nanofluid for each flooding test was 3–4 amphiphilic nanosheets was rather close to that of GO, implying
times the pore volume (PV). that no agglomeration took place after functionalization. From the
captured video of nanofluid dispersion, we did not find any no-
Results and Discussion ticeable aggregation. Photographs of amphiphilic nanosheet nano-
Surface Functionalization of GO with Alkylamine. The thickness of fluid in three locations at different time points were selected as
GO was estimated to be about 1 nm from the AFM image (Fig. S2), examples (Fig. 3C). Each bright dot in the photographs represented
indicating the single-layer feature of GO, in agreement with the a single nanosheet in Brownian motion. No aggregation, which
previous report (18). For comparison, we have also studied the would display as clusters, was detected. The excellent stability of the
effect of unmodified GO. Either the unmodified GO or our am- nanofluid would ensure that no extra additives or methods are re-
phiphilic Janus nanosheets were injected into the heptane/brine quired to preserve the nanofluid before injection.
system. The amphiphilic Janus nanosheets spontaneously accumu-
lated at the heptane/brine interface while GO stayed only in the Oil Displacement Efficiency and Mechanisms. Approximately 90% of
brine phase and instantaneously started to aggregate. When sub- the oil recovered by nanofluid flooding was extracted after the first
jected to vortex-induced vibrations, the amphiphilic Janus nano- PV injection. As shown in Table 1, under similar conditions (0.01%
sheets formed a thin interfacial film separating the heptane and nanofluid concentration), our recovery was 15.2%, more than triple
brine in contrast to the GO agglomeration due to the salt screening the best reported result of 4.7% (2). In addition, for 0.005 wt %

Fig. 2. (A) FTIR spectra and (B) TGA curves of GO and amphiphilic nanosheet.

7712 | www.pnas.org/cgi/doi/10.1073/pnas.1608135113 Luo et al.


Fig. 3. Hydrodynamic diameter distributions of (A) GO and (B) amphiphilic nanosheets in nanofluids with a concentration of 0.005 wt %. (Insets) Mode sizes. (C)
Selected photographs of amphiphilic nanosheets in the nanofluid in three locations at different time points.

concentration, which to our knowledge has not been previously study the oil displacement mechanisms, and our findings showed
reported, we also achieved extraordinary performance. results quite different from those of the traditional nanofluid.

APPLIED PHYSICAL
In reservoirs, the motion of crude oil can be categorized into As shown in Fig. 4, under static conditions, the interface be-

SCIENCES
three scenarios depending on the local underground hydrodynamic tween heptane and brine was concave in a water-wet glass tube but
power: static, slow, and fast moving. Therefore, we investigated the convex in an oil-wet plastic tube, indicating relatively strong in-
behaviors of nanosheets under simulated hydrodynamic scenarios to terfacial tension. When the nanofluid was injected into the brine,

Fig. 4. Behaviors of nanosheets in oil/brine system with increasing hydrodynamic power. (A–E) Behaviors of nanosheets in oil/brine system with increasing hy-
drodynamic power in a glass tube with water-wet surface (the heptane was dyed with Sudan Red 7B). (A) Heptane/brine mixture. (B) Nanosheets adsorbing to
heptane/brine interface during nanofluid injection under static conditions. (C) Nanosheets at heptane/brine interface after injection; a climbing film appeared on
the wall. (D) Growth of climbing film after gentle shaking. (E) Formation of interfacial film after vigorous shaking. (F–J) Behaviors of nanosheets in oil/brine system
with increasing hydrodynamic power in a plastic tube with oil-wet surface (Sudan Red 7B was not used as the whole plastic tube would be dyed). (F) Heptane/brine
mixture. (G) Nanosheets adsorbing to heptane/brine interface during nanofluid injection under static conditions. (H) Nanosheets at heptane/brine interface after
injection; a climbing film appeared on the wall. (I) Growth of climbing film after gentle shaking. (J) Formation of interfacial film after vigorous shaking.

Luo et al. PNAS | July 12, 2016 | vol. 113 | no. 28 | 7713
Table 1. Oil recovery factor of nanofluid flooding with different concentrations
Oil recovery factor Enhanced oil recovery
Average liquid Nanofluid after brine water factor after nanofluid
Rock core Porosity, % permeability, mD concentration, wt % flooding, % flooding, % Total oil recovery factor, %

1 24.8 54.4 0.005 71.1 6.7 77.8


2 26.0 44.5 0.01 62.5 9.5 72.0
3 27.9 130.0 0.005 68.2 10.2 78.4
4 25.8 132.0 0.01 69.6 15.2 84.8

electrostatic repulsion was screened by salt and hydrophobic at- side facing the water phase. Such behaviors altered the wettability
traction came into play (19). As a result, nanosheets spontane- of the oil-wet surface.
ously accumulated at the interface in each case. With increasing To simulate the condition of moderate hydrodynamic power,
amount of nanosheets adsorbed at the interface, the interfacial the tubes were shaken gently after injection. It was observed
tension was further reduced (20, 21), as indicated by the in- that for water-wet glass tubes, the climbing film grew to en-
creasingly flattened interface (Movie S1). As with conventional capsulate the oil phase and may carry the oil forward at flow
nanofluids, however, salt ions in the reservoir fluid were found to conditions, leaving very little residual oil behind. For the oil-
be a permeability damage factor due to increased nanoparticle wet plastic tube, the growth of the climbing film was not as ob-
aggregation. The climbing film was also observed because locally vious as with the glass tube, possibly due to the adsorption of
raised nanosheet concentrations induced Marangoni stress (the nanosheets onto the wall of the tube. When subjected to vigorous
stress produced by the gradient of interfacial tension) to push the shaking, the nanosheets formed flat films at interfaces for both the
oil–water interface up the tube surface (22). This film climbed glass and plastic tubes. The formation of such a solid-like film was
upward in the case of the water-wet tube surface, which helped
detach the oil phase at solid surface, but downward in the case of
the oil-wet tube surface. For the oil-wet surface, nanosheets may
be captured by the hydrophobic tube surface with the hydrophilic

Fig. 6. Testing the stability of interfacial films. (A and B) Deformation of in-


terfacial films upon tilting. (A) Image of heptane/brine interfaces in tilted glass
tubes with water-wet surfaces. (Upper) Interface with nanosheet interfacial film.
(Lower) Interface without nanosheets. (B) Image of heptane/brine interfaces in
Fig. 5. Testing the elasticity of interfacial films. (A) Interfacial film responding tilted plastic tubes with oil-wet surfaces. (Upper) Interface with nanosheet in-
to intrusion of a glass rod in a glass tube with a water-wet surface (from left to terfacial film. (Lower) Interface without nanosheets. (C and D) Reformation of
right: before intrusion, during intrusion, after intrusion). (B) Interfacial film interfacial films after shaking. (C) Image of interfacial film under vigorous
responding to intrusion of a glass rod in a plastic tube with an oil-wet surface shaking in a glass tube with water-wet surface (from left to right: before
(from left to right: before intrusion, during intrusion, after intrusion). (C) The shaking, during shaking, 1 min after shaking). (D) Image of interfacial film
heptane/brine interface with amphiphilic nanosheets (Left), SDS (Middle), and under vigorous shaking in a plastic tube with oil-wet surface (from left to
TWEEN 20 (Right) responding to intrusion of a glass rod. right: before shaking, right after shaking, 1 min after shaking).

7714 | www.pnas.org/cgi/doi/10.1073/pnas.1608135113 Luo et al.


Fig. 7. Schematic illustration of oil displacement mechanisms. (A) Climbing film encapsulation mechanism for water-wet surface. (B) Slug-like displacement mechanism.

also predicted by computer simulation for near-neutral wetting deformation but simply broke through, very different from the elastic
spherical particles (23). interfacial film formed by amphiphilic nanosheets (Fig. 5C). In

APPLIED PHYSICAL
When subjected to intrusion of a glass rod, the surface of in- terms of interfacial rheology, the interfacial films may resist dilation
terfacial films was deformed but not ruptured (Fig. 5 and Movie S2). and bending as characterized by the nearly flat interface after the

SCIENCES
After removal of the glass rod, the films recovered to their original tubes were tilted to enlarge the area of the interfaces (Fig. 6 A and B).
state, clearly demonstrating their elasticity. The presence of nano- Vigorous agitation disrupted the films for both solid surfaces (Fig. 6 C
sheets altered not only the normal stress balance, but also the tan- and D). However, the films reformed immediately at the interfaces
gential stress balance, leading to a redistribution of nanosheets to and separated the oil and brine phases after shaking. This process may
form a flat shape. The elasticity of the interfacial films kept them be driven by the amphiphilicity of nanosheets. In contrast to emulsion
intact even at 90 °C. To demonstrate that the elasticity of the inter- flooding, the interfacial film in our case strictly separates the water
facial film was unique to our amphiphilic nanosheets, we chose so- and oil phases, which may push the oil to the outlet like a slug at flow
dium dodecyl sulfate (SDS) and TWEEN 20 (purchased from conditions. This was confirmed in the core flooding tests, where the oil
and brine came out sequentially at the outlet. Such oil displacement
Sigma-Aldrich) as examples of ionic and nonionic surfactants, re-
mechanism leaves less oil residue and is free of demulsification, which
spectively, for comparison. With the same concentration of 0.01 wt %,
is a significant advantage of our nanofluid over the traditional ones.
when the interfaces formed by SDS or TWEEN 20 were subjected to After careful examination of the existing mechanisms (6–13), we
the glass rod intrusion, neither of them exhibited observable elastic found none of them was fully applicable to the above experimental
observation under different simulated hydrodynamic conditions.
Therefore, we propose two oil displacement mechanisms for
nanofluid flooding with nanosheets, as illustrated in Fig. 7:
i) The climbing film (a film of nanosheets along the tube’s surface)
encapsulation for water-wet surface. As shown in Fig. 7A, at
t = t0 the increased concentration of nanosheets due to the
adsorption at the oil–water interface produces the concentration
gradient leading to transfer of nanosheets to the three-phase
(nanofluid, oil, and rock solid) region, detaching and encapsu-
lating oil from the rock surface. When flow continues under
gentle hydrodynamic condition from t0 to t0 + Δt, the film grows
due to the ongoing supply of nanosheets from the nanofluid and
carries the oil phase forward.
ii) Slug-like displacement by the interfacial film. As seen in Fig. 7B,
at t = t0 an elastic interfacial film forms at the oil–water interface
at strong hydrodynamic power condition. The film can resist
bending and also reform after being disrupted. As a result, at
t = t0 + Δt oil is slug-like and displaced over a certain distance.
Except for these oil displacement mechanisms, the amphi-
philic Janus nanosheets are expected to have a lower chance to
be captured by the rock surface or to plug the rock pores due to
the self-accumulating at the oil–water interface, which may also
contribute to the high efficiency of oil recovery. The cross-sec-
tions of rock cores were examined by SEM before and after core
flooding and are displayed in Fig. 8. In each figure, the smallest
Fig. 8. SEM images of the cross-sections of (A) rock core with high per- and largest pore openings are both labeled. As for the rock cores
meability and (B) rock core with low permeability before core flooding tests; with high permeability (Fig. 8 A and C), after nanofluid flooding,
and (C) rock core with high permeability and (D) rock core with low per- the pore sizes at the two ends remained at the same level,
meability after core flooding tests. Arrows in each image indicate the comparable to those before flooding. Similar results were also
smallest and largest pore openings. detected in rock cores with low permeability (Fig. 8 B and D),

Luo et al. PNAS | July 12, 2016 | vol. 113 | no. 28 | 7715
with survival of narrow pore openings of around 2 μm. The with graphene-based Janus amphiphilic nanosheets at a low
observation that there were no noticeable changes in the pore concentration. The result from core flooding measurements
opening sizes indicated that our nanofluid caused minimal showed that the oil enhancement efficiency of 15.2% by this
damage to the permeability of the rock pores. nanofluid flooding is more than three times that of the pre-
In addition, fresh water was used to prepare the nanofluid for the viously reported best efficiency (4.7%) under similar conditions
tests described above. However, brine is preferred in some opera- at 0.01 wt % concentration. The behavior tests of nanosheets in
tions (e.g., seawater in offshore reservoirs or when fresh water is oil and brine system provided evidence that under a saline
scarce) to reduce cost and conserve fresh-water resources. This environment, (i) the accumulation of nanosheets at the oil–
requires that nanosheets be stable in a saline environment for a water interface, (ii) the appearance of climbing films, and (iii) the
certain time before reaching underground reservoirs. Our study generation of elastic interfacial films may be responsible for the
showed that sequential addition of polyvinylpyrrolidone and poly- high oil recovery efficiency.
vinyl alcohol provided good stability to our nanofluid in brine. We
are currently investigating the stabilization mechanism, the oil re- ACKNOWLEDGMENTS. We thank Ishwar Mishra and Prof. Dong Cai in the
covery efficiency in core flooding tests, and oil displacement Department of Physics, University of Houston, for providing particle size
mechanisms due to the existing polymer. analysis on Malvern NS300. The work performed at University of Houston is
supported in part by the US Department of Energy under Contract DE-FG02-
Conclusion 13ER46917/DE-SC0010831, US Air Force Office of Scientific Research Grant
FA9550-15-1-0236, the T.L.L. Temple Foundation, the John J. and Rebecca
We have presented the first (to our knowledge) tertiary or en- Moores Endowment, and the State of Texas through the Texas Center for
hanced oil recovery experiments using simple nanofluid flooding Superconductivity at the University of Houston.

1. Torsater O, Engeset B, Hendraningrat L, Suwarno S (2012) Improved Oil Recovery by 12. Kondiparty K, Nikolov AD, Wasan D, Liu KL (2012) Dynamic spreading of nanofluids
Nanofluids Flooding: An Experimental Study. (Society of Petroleum Engineers, on solids. Part I: Experimental. Langmuir 28(41):14618–14623.
Richardson, TX). 13. Liu KL, Kondiparty K, Nikolov AD, Wasan D (2012) Dynamic spreading of nanofluids
2. Hendraningrat L, Li SD, Torsaeter O (2013) A coreflood investigation of nanofluid on solids. Part II: Modeling. Langmuir 28(47):16274–16284.
enhanced oil recovery. J Petrol Sci Eng 111:128–138. 14. Shen PP, Wang JL, Yuan SY, Zhong TX, Jia X (2009) Study of enhanced-oil-recovery
3. Torsaeter O, Li S, Hendraningrat L (2013) Enhancing Oil Recovery of Low-Permeability mechanism of alkali/surfactant/polymer flooding in porous media from experiments.
Berea Sandstone Through Optimised Nanofluids Concentration. (Society of Petroleum Soc Pet Eng J 14(2):237–244.
Engineers, Richardson, TX). 15. Marcano DC, et al. (2010) Improved synthesis of graphene oxide. ACS Nano 4(8):
4. Hendraningrat L, Torsaeter O (2015) Metal oxide-based nanoparticles: Revealing their 4806–4814.
potential to enhance oil recovery in different wettability systems. Appl Nanosci 5(2): 16. Hong L, Jiang S, Granick S (2006) Simple method to produce Janus colloidal particles
181–199. in large quantity. Langmuir 22(23):9495–9499.
5. Zargartalebi M, Kharrat R, Barati N (2015) Enhancement of surfactant flooding per- 17. Wu H, Yi WY, Chen Z, Wang HT, Du QG (2015) Janus graphene oxide nanosheets
formance by the use of silica nanoparticles. Fuel 143:21–27. prepared via Pickering emulsion template. Carbon 93:473–483.
6. Binks BP (2002) Particles as surfactants - similarities and differences. Curr Opin Colloid 18. Dikin DA, et al. (2007) Preparation and characterization of graphene oxide paper.
Interface Sci 7(1-2):21–41. Nature 448(7152):457–460.
7. Zhang H, Nikolov A, Wasan D (2014) Enhanced oil recovery (EOR) using nanoparticle 19. Chen Q, Bae SC, Granick S (2011) Directed self-assembly of a colloidal kagome lattice.
dispersions: Underlying mechanism and imbibition experiments. Energy Fuels 28(5): Nature 469(7330):381–384.
3002–3009. 20. Binks BP, Clint JH, Fletcher PDI, Lees TJG, Taylor P (2006) Growth of gold nano-
8. Giraldo J, Benjumea P, Lopera S, Cortes FB, Ruiz MA (2013) Wettability alteration of particle films driven by the coalescence of particle-stabilized emulsion drops.
sandstone cores by alumina-based nanofluids. Energy Fuels 27(7):3659–3665. Langmuir 22(9):4100–4103.
9. Karimi A, et al. (2012) Wettability alteration in carbonates using zirconium oxide 21. Cui M, Emrick T, Russell TP (2013) Stabilizing liquid drops in nonequilibrium shapes by
nanofluids: EOR implications. Energy Fuels 26(2):1028–1036. the interfacial jamming of nanoparticles. Science 342(6157):460–463.
10. Lim S, Horiuchi H, Nikolov AD, Wasan D (2015) Nanofluids alter the surface wettability 22. Cheng HL, Velankar SS (2008) Film climbing of particle-laden interfaces. Colloid
of solids. Langmuir 31(21):5827–5835. Surface A 315(1-3):275–284.
11. Wasan DT, Nikolov AD (2003) Spreading of nanofluids on solids. Nature 423(6936): 23. Stratford K, Adhikari R, Pagonabarraga I, Desplat JC, Cates ME (2005) Colloidal jamming
156–159. at interfaces: A route to fluid-bicontinuous gels. Science 309(5744):2198–2201.

7716 | www.pnas.org/cgi/doi/10.1073/pnas.1608135113 Luo et al.

You might also like