Synergy - Between - Nanoparticles - and - Surfactants - in - Stabilizing - Foams - For - Oil - Recovery

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

Article

pubs.acs.org/EF

Synergy between Nanoparticles and Surfactants in Stabilizing Foams


for Oil Recovery
Robin Singh and Kishore K. Mohanty*
Department of Petroleum and Geosystems Engineering, The University of Texas at Austin, Austin, Texas 78712, United States

ABSTRACT: Surfactant-stabilized foams have been used in the past for vertical conformance and mobility control in gas-
enhanced oil recovery processes. Lack of stability of these foams often limits their application. The goal of this study is to
investigate the synergistic effects of using a blend of silica nanoparticles (NPs) and anionic surfactants on foam stability in both
bulk and porous media. First, the stability of static foams was studied using surfactants and surfactant−NP mixtures with and
without the presence of a crude oil. Second, the foam drainage behavior and thickness of the foam lamella were studied by
fluorescence microscopy. Third, the mobility of foams was measured by coinjecting the surfactant or surfactant−NP solution
with nitrogen gas through a Berea sandstone core at a fixed foam quality (gas volume fraction). Finally, oil displacement
experiments were conducted in Berea cores using these foams. Static foam tests indicated a stabilization effect of nanoparticles on
surfactant−NP-stabilized foams in the absence of crude oil. Adding nanoparticles at low concentrations (0.3 wt %) improves the
foam stability and increases the mobility reduction factor by a factor of 2 in the absence of oil. Fluorescence and confocal laser
scanning microscopy elucidated the trapping of nanoparticles in Plateau borders as well as lamellas, which retards liquid drainage
and bubble coalescence. Core floods with a reservoir crude oil showed about 10% incremental oil recovery by an immiscible foam
(with the surfactant−NP blend) over a water flood. This study shows that nanoparticles have the potential to increase the
stability of surfactant-stabilized foams in subsurface applications.

■ INTRODUCTION
Gas flooding is one of the most widely used enhanced oil
essential mechanism of foam transport.13 The economics of
surfactant-stabilized foam depends on the quantity of surfactant
recovery (EOR) techniques.1 It consists of injection of a gas required for the long -distance propagation of foam from the
(e.g., hydrocarbon components such as methane, propane, and wellbore. Various factors such as adsorption, surfactant loss due
enriched gases and non-hydrocarbon components such as to partitioning into crude oil, and surfactant degradation under
carbon dioxide and nitrogen) into previously water-flooded oil harsh reservoir conditions further limit the economic viability
reservoirs. In the United States, CO2 EOR projects alone of surfactant usage in subsurface applications.14
provide 280 000 barrels of oil per day, which is 3% of domestic The use of nanoparticles (NPs) can help mitigate these
oil production.2 If the gas is first-contact- or multicontact- issues. Nanoparticles have the potential to stabilize foam under
miscible with the oil, it can displace virtually all of the oil in the harsh conditions of temperature and salinity. Moreover, these
volume swept by the gas.3 However, the adverse mobility ratio nanoparticles can be obtained cost-effectively from cheap raw
due to the inherent low viscosity of gas leads to viscous materials such as fly ash and silica.15 Many studies have
fingering, leaving a part of the reservoir uncontacted.4 Reservoir reported the use of colloidal particles with different surface
heterogeneity and gravity override also contribute to poor chemistry as sole stabilizers for bulk foam.16−18 Most of these
sweep efficiency.5 particles were coated with surfactants or polymers to act as
Foam can be used to mitigate poor sweep problems surface-active agents, as they were not amphiphilic themselves.
associated with gas injection.6,7 The main mechanism by Alargova et al.16 achieved superstabilized aqueous foams using
which foam reduces the gas mobility is by immobilizing or synthesized polymer microrods in the absence of surfactants.
trapping a large fraction of the gas in the porous medium and These foam bubbles were sterically stabilized by entangled rod-
by increasing the apparent viscosity of the gas.8 The concept of shaped structures of the polymer. Binks and Horozov17
using foams as mobility control agents was first proposed by reported aqueous foams stabilized solely by silica nanoparticles
Bond and Holbrook9 in 1958. Since then, there have been with different degrees of hydrophobicity. These stable foams
several successful foam pilot tests in which surfactant and gases
were formed by adsorption of aggregates (few hundred nm) of
have been coinjected for mobility control.10 In bulk, foam is
nanoparticles at the surfaces. Another way to obtain surface-
defined as a dispersion of gas bubbles in liquid; the bubble size
is much smaller than the size of the container.11 The foam active particles is adsorption of amphiphiles on the particle
morphology in porous media is quite different than in the bulk; surface. Gonzenbach and co-workers18 employed this technique
the bubble size is close to the size of pores. In porous media, to generate stable macroscopic foams. The unique colloidal
foam is a two-phase fluid system where some gas flow paths are architecture, i.e., the sequential assembly of amphiphiles on the
made discontinuous by thin liquid films called lamellas.12 These
lamellas are typically stabilized by surfactants that are adsorbed Received: July 3, 2014
at the interface. However, lamellas lack long-term stability. In Revised: November 16, 2014
porous media, continuous regeneration of these lamellas is an Published: December 26, 2014

© 2014 American Chemical Society 467 DOI: 10.1021/ef5015007


Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

surface of the particle and particles to the air−water interface, measured. Water floods and subsequent foam floods were
was the proposed mechanism for long-term stability of foam. conducted in Berea cores saturated with a crude oil using
Espinosa et al.19 investigated supercritical CO2-in-water (C/ surfactant or a surfactant−nanoparticle mixture as the foaming
W) foam generation in bead packs using hydrophilic silica agent.
nanoparticles coated with poly(ethylene glycol) (PEG). The
generated foam using nanoparticles had 2−18 times more
resistance to flow than the same fluid without nanoparticles. Yu
■ METHODOLOGY
Materials. The anionic surfactant Bioterge AS-40, a C14−16 α-olefin
et al.20 studied the effects of the CO2−nanoparticle dispersion sulfonate (39% active), was obtained from Stepan Co. The
phase ratio and total flow rate on CO2 foam mobility (the ratio nanoparticles used in this study were aqueous dispersions of silica
of core effective permeability and foam viscosity) in bead packs. nanoparticles, Nyacol DP 9711, as provided by Nyacol Nano
The nanoparticle-stabilized foam mobility increased with Technologies, Inc. Fluorescent silica nanoparticles (FL-NPs), obtained
from 3M, were used for fluorescence microscopy. The primary silica
increases in both the dispersion phase ratio and the total flow
particles were 5 nm in diameter, and the PEG coating brought the
rate. Mo et al.21 continued this study to understand the CO2 particles to a nominal diameter of 10 nm. Berea sandstone cores were
foam behavior in Berea sandstone cores. They reported foam used in foam flow, retention, and core flood experiments. Crude oils
generation with nanoparticle concentrations as low as 100 ppm. (A, B, and C) were obtained from a reservoir. The properties of these
The foam mobility attained a minimum when the foam quality crude oils are tabulated in Table 1. The viscosity was measured using
was between 40% and 80%. Aroonsri et al.22 investigated C/W an ARES rheometer. Sodium chloride (>99% pure, Fisher Chemical)
foam generation through bead packs and fractured and and nitrogen (research grade, Matheson) were used as received.
unfractured sandstone cores using various silica nanoparticles.
They emphasized the role of proper hydrophilic−CO2-philic Table 1. Properties of the Crude Oils
balance (HCB) of nanoparticle selection for viscous C/W foam
crude density (g/ viscosity (cp) at acid number (mg of KOH/g
generation. Foam flow experiments in cores showed that a oil cm3) 25 °C of oil)
critical shear rate exists for foam generation. Worthen et al.23 A 0.773 9 2.45
demonstrated the generation of viscous C/W foams in bead B 0.887 382 1.72
packs using partially hydrophobic silica nanoparticles and C 0.830 36.2 4.91
showed that these foams were more stable than foams
generated using hydrophilic PEG-coated silica nanoparticles. Aqueous Stability of Nanoparticles. The Nyacol DP 9711 silica
Surface modification of nanoparticles via chemical treatment nanoparticles have proprietary coating that makes them hydrophilic in
can sometimes be expensive. In such cases, altering the surface nature. The size of the nanoparticles was characterized by transmission
properties of the particles by in situ physioadsorption of electron microscopy (TEM). A droplet of the nanoparticle dispersion
surfactants on their surfaces could be cost-effective. Several was placed on a Formvar-coated copper grid and analyzed using a FEI
researchers have investigated the potential of utilizing this Tecnai transmission electron microscope operating at 80 kV. The
mean diameter of the primary particles was found to be 20 nm via
synergism between surfactants and nanoparticles to generate
image analysis using ImageJ software. The TEM image is shown in
foam.24−27 Zhang et al.25 demonstrated the synergistic Figure 1. The particle hydrodynamic diameters in the aqueous
stabilization of aqueous foams by mixtures of hydrophilic
Laponite particles and a nonionic surfactant, tetraethylene
glycol monododecyl ether, at surfactant concentrations higher
than 2 wt %. Adsorption of the modified particles on the air−
water interface was confirmed by laser-induced confocal
microscopy. Cui et al.27 showed that non-surface-active
CaCO3 nanoparticles can be surface-activated via interaction
with anionic surfactants, leading to enhanced foam stability.
Recently, Worthen et al.28 reported fine-textured C/W foam
generation in bead packs using the synergistic interaction of
hydrophilic silica nanoparticles and a surfactant, caprylamido-
propyl betaine, when neither of these species could stabilize
foam independently.
The interactions of surfactants and nanoparticles and their Figure 1. TEM image of surface-modified silica nanoparticles (scale
bar is 100 nm).
effect on foam transport in porous media are not well
understood. The objective of this paper is to systematically
study synergistic stabilization of foams by mixtures of surface- dispersions were characterized via dynamic light scattering (DLS)
modified nanoparticles and anionic surfactants in bulk and using a DelsaNano analyzer without any pH adjustment at room
porous media with and without the presence of crude oil. The temperature. Ultrapure water with a resistivity greater than 18.2
sensitivity of foams to various crude oils was studied and was MΩ-cm was used to prepare brine solutions. The stock solution of 2
correlated to parameters such as lamella number and spreading, wt % nanoparticles was first prepared. This stock solution was used to
entering, and bridging coefficients. Vertical foam films were prepare six samples containing various concentrations of NaCl (1, 2, 4,
inspected using fluorescently tagged nanoparticles to study the 6, 8, and 10 wt %) and 0.5 wt % nanoparticles. The seventh sample
foam drainage behavior. The mechanism of foam stability was had 0.5 wt % nanoparticles in API brine (8 wt % NaCl and 2 wt %
CaCl2). Figure 2 shows these samples after 2 weeks of preparation.
analyzed in terms of detachment energy, maximum capillary The sample solutions were clear, and no signs of any precipitation
pressure of coalescence, and stratification. Foams were created were observed. Table 2 shows the measured hydrodynamic diameters
in situ by coinjecting surfactant−nanoparticle mixtures and of the nanoparticles for these samples. The mean values were between
nitrogen gas through a Berea sandstone at a fixed quality 24.9 and 35.5 nm for different salinities, indicating no particle
(volume fraction of gas), and mobility reduction factors were aggregation. These values from DLS analysis are slightly larger than

468 DOI: 10.1021/ef5015007


Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

evaporation. The height of the foam above the liquid phase was
monitored as the time progressed. In foam tests in the presence of
crude oils, first the foam was generated by dispersing the air and then
oil (2 mL) was introduced from the bottom using a very fine tube.
Vertical Foam Film Tests. Zhang33 performed this test to study
the effect of hardness on foam stability. A similar method was adopted
in this study to examine the effect of nanoparticles on foam stability.
The foam drainage behavior was visualized by making foam in a small,
vertically oriented optical glass cell (4 mm × 12 mm × 48 mm;
Produstrial, Fredon, NJ). A small amount of nanoparticle dispersion
Figure 2. Samples containing 0.5 wt % NPs and (from left to right) 1, (0.5 mL) was injected into the cell, and foam was generated by
2, 4, 6, 8, or 10 wt % NaCl or API brine. dispersing air from the bottom using a syringe. The time-dependent
foam morphology was observed using a Nikon microscope equipped
Table 2. Diameters of Nanoparticles Measured Using DLS with a high-resolution camera. Images were recorded every 30 s using
Image Pro software, and the lamella width was measured after
sample diameter (nm) sufficient drainage time. Similar experiments were conducted using the
0.5 wt % NPs in 1 wt % NaCl 28.6 ± 7.7 dispersion of fluorescently tagged nanoparticles. Images of foam
0.5 wt % NPs in 2 wt % NaCl 24.9 ± 6.9 texture were captured in both visible and UV light. All of these
0.5 wt % NPs in 4 wt % NaCl 28 ± 7.4 measurements were done at room temperature.
0.5 wt % NPs in 6 wt % NaCl 30.3 ± 8 Confocal Laser Scanning Microscopy (CLSM). CLSM is a
0.5 wt % NPs in 8 wt % NaCl 31.5 ± 8.1
powerful tool that has often been used in the literature34,35 to study
nanoparticles in foam films. CLSM was performed with a Leica SP2
0.5 wt % NPs in 10 wt % NaCl 29.1 ± 8.1
AOBS confocal microscope using a 10× dry (HC PL APO 0.40NA
0.5 wt % NPs in API brine 35.5 ± 8.8 CS) objective lens operating in fluorescence mode. Samples containing
0.5 wt % surfactant and 0.3 wt % fluorescently tagged nanoparticles
were used. The nanoparticles were excited with a 488 nm wavelength
that from TEM, as expected because DLS measures the hydrodynamic
laser. In order to visualize the low-quality foam (disconnected
diameter and is biased toward larger particles in the suspensions.29,30
bubbles), a small amount of sample was placed in the welled slide
Preparation of Nanoparticle Dispersions. A stock solution of 1
wt % Bioterge (BT) surfactant was first prepared. Samples were and air was dispersed using a syringe to make foam. A coverslip was
prepared by diluting this stock solution. Aqueous dispersion samples then placed on top of the slide to seal it. The bottom of the sample
were prepared by incorporating silica nanoparticles at concentrations was then scanned at room temperature. For imaging of the high-
of 0, 0.1, 0.3, and 0.5 wt % along with 0.5 wt % surfactant and 1 wt % quality foam (connected bubbles), a procedure similar to that for the
sodium chloride. The dispersions were stirred for 12 h to ensure vertical foam film tests was adopted. A small amount of sample was
mixture homogeneity. Table 3 shows the diameter and ζ potential placed in a rectangular glass cell (4 mm × 12 mm × 48 mm;
(both measured using the DelsaNano analyzer) of each nanoparticle Produstrial), and foam was generated by dispersing air using a syringe.
dispersion with and without surfactant. As we can see from this table, The glass cell was then sealed and placed horizontally, and the middle
the presence of surfactant does not cause any particle agglomeration, section of sample was scanned under the microscope. The imaging
suggesting that the BT surfactant can be used in combination with the resolution was 200 μm for both of the above cases.
nanoparticles. Foam Flow Experiments. The cores were dried at 90 °C for 24 h
Interfacial Surface Tension (IFT) Measurements. The IFTs in in an oven and laminated with FEP shrink-wrap tubing (Geophysical
oil−brine, oil−air, and brine−air systems were measured via pendant Supply Company, Houston, TX). These were then placed in a Hassler-
droplet analysis using a Ramé-Hart goniometer. The droplet (of oil or type core holder (Phoenix, Houston, TX) with a confining pressure of
water) was held for a sufficient time (>5 min) to allow it to equilibrate 1500 psi. The brine porosity and permeability of the cores were then
with the second phase (air or liquid). Axisymmetric shape analysis of determined using standard methods.36 Figure 3 shows a schematic of
the droplet was performed using DROPimage Advanced software, the experimental setup. Two series-D syringe pumps from Teledyne
which measures the IFT by fitting the drop profile with the solution to ISCO (Lincoln, NE) that are capable of low injection rates (as low as
the Young−Laplace equation. Ten measurements were performed 5 s 0.001 cm3/min) were used. The apparatus was built to coinject
apart; the means and standard deviations of these measurements were nitrogen gas and surfactant or a surfactant−NP blend through a
calculated. sandpack (diameter, 0.6 in.; length, 6 in.) to ensure proper mixing and
Foamability and Static Foam Tests. Preliminary foam tests were foam generation. The pregenerated foam was then injected from the
conducted by shaking 5 mL of these samples vigorously 10 times in top of the core in the core holder. The effluent from the core went to a
test tubes, and the macroscopic foam textures were observed at room high-pressure view cell. The downstream pressure of the experiment
temperature. In the case of foam shake tests in the presence of crude was maintained by a back-pressure regulator installed after the view
oil, 0.5 mL of oil was initially added to the sample before vigorous cell. The pressure drops across various sections of the core were
mixing. Static foam tests were conducted using a similar method as measured using Rosemount differential pressure transducers. All of the
reported in the literature.31,32 A 100 mL aliquot of each sample was connections were made with stainless steel Swagelok fittings. At the
then added to a graduated glass cylinder (diameter, 4 cm; length, 30 end of this experiment, several pore volumes (PV) of brine was flushed
cm). Air was injected from the bottom, generating the static foam. The through the core to remove the foam, and the back pressure was
cylinder was sealed from the top using a rubber stopper (with a small intermittently depressurized and pressurized several times to remove
outlet port for venting air during foam generation) to avoid any the trapped gas. Finally, the brine permeability was measured. A similar

Table 3. Properties of Nanoparticle Dispersions

sample diameter (nm) by DLS ζ potential (mV)


0.5 wt % NPs + 1 wt % NaCl 28.6 ± 7.7 −3.87 ± 2.56
0.5 wt % BT + 0.5 wt % NPs + 1 wt % NaCl 28.3 ± 8.2 −4.34 ± 1.93
0.5 wt % FL-NPs + 1 wt % NaCl 15.3 ± 6.6 −3.47 ± 2.86
0.5 wt % BT + 0.5 wt % FL-NPs + 1 wt % NaCl 16.1 ± 7.3 −3.98 ± 3.21

469 DOI: 10.1021/ef5015007


Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

Figure 3. Schematic of the apparatus for foam flow experiments and core floods.

procedure was adopted by Simjoo and co-workers37,38 during their Andrianov et al.40 in their study of immiscible foam for enhanced oil
studies of immiscible foam flow in porous media. recovery.
Nanoparticle Transport and Retention in Porous Media. The
experimental setup was designed to quantify the amount of
nanoparticle retention in the porous medium. The setup was similar
to that shown in Figure 3 except that the sandpack and view cell were
■ RESULTS AND DISCUSSION
Foamability and Foam Sensitivity to Crude Oil. Basic
removed. The silica nanoparticle concentration in the effluent was shake tests were performed using 0.5 wt % surfactant in 1 wt %
measured by inductively coupled plasma optical emission spectrometry NaCl brine at various nanoparticle concentrations from 0 to 0.5
(ICP-OES) following the method described in Wang et al.39 The wt %. The macroscopic foam behavior was observed. Figure 4
calibration curve for different nanoparticle dispersion standards was
found to be linear (R2 = 0.9998).
Oil Displacement Experiments. The experimental setup was
similar to that shown in Figure 3 except that the view cell was removed
to decrease the dead volume of the setup in order to accurately
measure the oil recovery. The cores were first fully saturated with
brine; the brine porosity and permeability were measured using
standard methods. These core properties are tabulated in Table 4. The

Table 4. Properties of the Cores Used


initial oil
length diameter porosity permeability saturation
experiment (cm) (cm) (%) (mD) (%)
foam flow 29.36 2.52 23.5 357 0
retention 30.27 2.50 22.3 313 0
core flood 1 30.30 2.47 21.3 383 64
core flood 2 30.30 2.49 22.9 315 63.8
Figure 4. Foam morphologies with 0.5 wt % surfactant at nanoparticle
cores were then flooded with filtered crude oil (at least 2.5 PV) from concentrations of (from left to right) 0, 0.1, 0.3, and 0.5 wt %.
the top at a constant pressure of 750 psi at room temperature until the
brine production stopped. The initial oil saturation was determined by
mass balance. The whole setup was pressurized with a back pressure of shows that these samples achieved the maximum foam height
100 psi. The brine flood was conducted at 1 ft/day for 2 PV until no possible in the vial, indicating foaming tendency in the absence
oil was produced. Then the cores were flooded with brine at 5 ft/day of oil. No visual differences were observed (by naked eye) in
to minimize capillary end effects. The cores were then preflushed with the macroscopic foam texture with increasing nanoparticle
1 PV of surfactant or surfactant−NP blend to avoid any adsorption of
surfactant during foam flooding. Nitrogen gas and surfactant or concentration in the surfactant solution. This experiment was
surfactant−NP blend were then coinjected through the sandpack to then repeated in the presence of crude oil. Before mixing of the
make foam. This pregenerated foam was then injected into the core solution, 0.5 mL of crude oil was added to the solution. Three
from the top for more than 7 PV. Oil recovery and pressure drops crude oils (A, B, and C) were used in this test. Figure 5 shows
were monitored at each step. Similar methodology was adopted by the foam morphologies at time t = 0 min formed using 0.5 wt %
470 DOI: 10.1021/ef5015007
Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

Figure 5. Foam morphologies with 0.5 wt % surfactant in the presence of oils A−C at (left) t = 0 min and (right) t = 8 h.

surfactant solution with no nanoparticles in the presence of to indicate whether oil can be imbibed into foam lamellas.
these crude oils. Foams have been classified into three types on the basis of the
It was seen that the foam volume decreased drastically in a L value: L < 1, 1 < L < 7, and L > 7 correspond to types A, B,
few minutes for the case of crude oil A, indicating that the foam and C, respectively. Type A foams represent the most stable
was not stable with this crude oil. Oil was instantaneously foams, which do not interact with oil. They have negative values
drained to the bulk solution, rupturing the foam lamellas, and for both E and S. Type B foams have negative S values and
no retention ability of oil within a lamella was observed. For positive E values and are moderately stable. Oil interacts with
crude oils B and C, the foam volume was retained at the foam lamellas but does not rupture them. Type C foams are
maximum at time t = 0 min and even for longer duration (t > 8 unstable, with positive values of both S and E. Oil tends to be
h), as shown in Figure 5. In both of these samples, oil was seen imbibed into these foams and rupture the foam lamellas.
distinctly inside the foam lamellas, indicating the tendency of Table 5 shows the measured interfacial surface tensions of
oil to be imbibed into the lamella without rupturing it. Thus, oil−gas−surfactant systems for the three crude oils. The
the foam was more stable even in the presence of crude oils B
and C compared with the case of crude oil A. This test was Table 5. Interfacial Tension Data for the Three Oils
repeated with samples having 0.5 wt % surfactant and 0.3 wt %
interfacial tension (mN/m)
nanoparticles, and the same macroscopic foam behavior was
observed. system oil A oil B oil C
It can be seen that foam behaves differently with different oil−gas 21.31 ± 0.06 22.16 ± 0.04 20.73 ± 0.06
crude oils. Typically, oils are known to be detrimental to foam oil−surfactant 0.08 ± 0.02 4.03 ± 0.01 2.35 ± 0.04
stability and are generally considered as foam-inhibiting or air−surfactant 31.69 ± 0.09 31.69 ± 0.09 31.69 ± 0.09
antifoaming agents.41 Many researchers have reported this
destabilizing effect of oil on foam stability in porous
media.42−44 Although the phenomenon of foam−oil inter- presence of nanoparticles (used in this study) did not alter the
actions is quite complex, a qualitative analysis can be done by surface tension values significantly, so the values of the
investigating several parameters that are dependent on the parameters S, E, L, and B would be the same for the case of
surface energy of the system. Three such parameters are the the oil−gas−surfactant−NP system. Figure 5 shows that after 8
spreading coefficient (S), the entering coefficient (E), and the h the foams were not stable with crude oil A (μ = 9 cP) and
bridging coefficient (B), which are defined as follows: were quite stable with crude oils B (μ = 382 cP) and C (μ =
36.2 cP). Thus, the sensitivities of the foam to the crude oils
S = σw/g − σw/o − σo/g (1) revealed that lighter crude oils are more destabilizing to foams
than heavier oils. The calculated values of the spreading
E = σw/g + σw/o − σo/g (2) coefficients, entering coefficients, lamella numbers, and bridging
coefficients are listed in Table 6. The results show that S, E, and
B = σw/g 2 + σw/o 2 − σo/g 2 (3) B were positive for all three crude oil systems. Thus, these
where σ is the interfacial tension and the subscripts w, g, and o parameters were not reliable in predicting the stability of the
correspond to water, gas, and oil, respectively. From foam−oil system. The lamella numbers predicted system with
thermodynamics, a negative value of S implies that foam will
be stable in the presence of oil.45 Similarly, the foam−oil system Table 6. Calculated Parameter Values Based on Interfacial
is stable when E is negative.46 A positive value of B is a Data
necessary (but not sufficient) condition for the oil to act as an
antifoaming agent.41 Schramm and Novosad47 have suggested parameter oil A oil B oil C
the parameter known as the lamella number (L), defined as S 10.3 ± 0.17 5.4 ± 0.14 8.61 ± 0.19
E 10.4 ± 0.17 13.5 ± 0.14 13.31 ± 0.19
σw/g
L = 0.15 L 59.4 ± 0.11 1.2 ± 0.10 2.02 ± 0.13
σw/o (4) B 550.1 ± 0.34 529.2 ± 0.28 579.9 ± 0.38

471 DOI: 10.1021/ef5015007


Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

crude oil A to be of type C and the systems with crude oils B


and C to be of type B, which is consistent with the macroscopic
observations of these foams. However, the criterion based on
lamella number implies that a low oil−surfactant water tension
system cannot form stable foams, which is not true.48
Static Foam Tests. Static foam tests were carried out with
nanoparticles dispersed in 0.5 wt % surfactant at concentrations
of 0.1−0.5 wt %, and the foam height decay was compared with
that of 0.5 wt % surfactant alone. The decay of the foam height
was monitored over time, and the results are plotted in Figure
6. The half-life, defined as the time required for the foam to

Figure 7. Static foam tests with crude oil A at 25 °C.

of crude oil A was on the order of only minutes (24 min).


Crude oils B and C had no detrimental effect on the foam. The
oil did not rupture the lamellas on contact, which resulted in
longer half-lives. Additional stability of the foam was observed
in surfactant−nanoparticle blends compared with surfactant
alone, as evidenced by the longer half-lives. The half-life
increased from 15 to 22 h for the case of crude oil B (Figure 8)
and from 10 to 14 h for the case of crude oil C (Figure 9).

Figure 6. Static foam tests with 0.5 wt % surfactant and various


nanoparticle concentrations at 25 °C.

decay to half of its original height, can be seen from the plot.
The half-life of foams generated using surfactants alone without
nanoparticles was about 48 h. As the concentration of
nanoparticles was increased to 0.1 wt %, the half-life increased
to 68 h. The synergistic effects on foaming of the surfactant−
nanoparticle system became more pronounced when the
nanoparticle concentration was increased to more than 0.3 wt
%. The foam heights remained almost constant for >4 days.
The mechanisms behind this synergy and enhanced bulk foam
stability resulting from the combination of surfactant and
nanoparticles were studied as described in the subsequent
sections.
These foam tests were then repeated in the presence of crude
Figure 8. Static foam tests with crude oil B at 25 °C.
oil A, B, or C using 0.5 wt % surfactant with or without
nanoparticles at room temperature. The nanoparticle concen-
tration used in these cases was 0.3 wt %. The foam was first Vertical Foam Film Tests. Foam was generated in the
generated, and then 2 mL of crude oil was introduced from the optical glass cell using 0.5 wt % surfactant solution with or
bottom using a fine tube. Macroscopic foam−oil interactions without nanoparticles. The concentration of nanoparticles in
and the foam height were observed with time. Figure 7 shows the former case was 0.3 wt %. The bubbles were formed
the foam height decay profile in the presence of crude oil A as between the parallel walls of foam cells spanning both sides.
well as foam−oil interactions at the oil interface. The image was The typical bubble structure can be considered as polygonal
captured just after the introduction of oil into the foam system. prisms. A Gibbs−Plateau border is formed when three
As seen in the preliminary foam tests, crude oil A was quite neighboring bubbles meet at one edge. The foam morphology
detrimental to the foam. As soon as the oil came into contact was observed using a high-resolution microscope. After a
with the foam, it resulted in rupturing of the lamellas, causing a sufficient drainage time of 30 min, images of the foam were
rapid decay of the foam with time. The half-life in this case was captured. Figure 10 shows the foam structures for the surfactant
only 24 min for both surfactant and surfactant−NP blend. No and surfactant−NP blend. The typical lamella width in the
additional stabilization of foam was observed due to presence of surfactant-only case was found to be about 340 μm, while for
nanoparticles, as the deleterious effect of the oil on the foam the blend it was 472 μm. The experiment was repeated several
dominated the foam decay. The half-life in the absence of crude times to check for reproducibility, and the results were
oil was on the order of days (>48 h), and that in the presence consistent with a variation of ±20 μm. The greater lamella
472 DOI: 10.1021/ef5015007
Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

using the same FL-NPs. Figure 12A−C shows fluorescence,


bright-field, and merged images, respectively, of a weak foam
sample. Since the nanoparticle concentration was relatively high
(0.3 wt %), the NPs (green) could be seen surrounding the
bubbles and dispersed in the continuous liquid phase. Figure
12D−F shows fluorescence, bright-field, and merged images,
respectively, of a strong foam sample. Silica nanoparticles
(green) could be observed surrounding the air−water interface
as well as the Plateau border. These nanoparticles form three-
dimensional networks that enhance the stability of the bubbles.
Mechanisms for Foam Stability in Bulk. The synergistic
stabilization of foams by surfactant−nanoparticle systems
depends on the interplay of nanoparticle−interface, surfac-
tant−interface, and nanoparticle−surfactant interactions. Some
of the key mechanisms presented in the literature for particle-
stabilized foams are discussed below for surfactant−nano-
particle systems.
Figure 9. Static foam tests with crude oil C at 25 °C. Particle Detachment Energy. The energy required to detach
a particle of radius R from the interface depends on the contact
angle (θ) and the surface tension of the interface (γaw).49 If the
particle size is quite small (less than a few microns), gravity and
buoyancy effects can be neglected. The amount of energy
required to move the particle from the interface to the bulk
solution is given by
Ed = πR2γaw(1 − |cos θ|)2 (5)

For θ < 30° (highly hydrophilic particle) or θ > 150° (highly


hydrophobic particle), this detachment energy would be quite
Figure 10. Vertical foam films made using (A) 0.5 wt % surfactant and low, which implies that these particles cannot stabilize foam (as
(B) 0.5 wt % surfactant + 0.3 wt % NPs (scale bars are 1000 μm).
per this theory). For a contact angle close to 90° and a particle
of diameter 10 nm, this detachment energy is quite large in the
width for the surfactant−NP blend case indicates a retarded
absence of surfactants, on the order of 103kBT. The large energy
drainage process.
associated with attachment implies that once the nanoparticle is
To fundamentally understand the mechanism of enhanced
brought to the interface, it is irreversibly adsorbed on the
foam stability and greater lamella width for the case of the
interface and provides robust foam stability.50 The Nyacol NPs
surfactant−NP blend, fluorescently tagged nanoparticles were
coated with a proprietary coating (20 nm) and the 3M FL-NPs
used. Foam was generated using 0.5 wt % surfactant with 0.3 wt
coated with PEG (10 nm) are 1 order of magnitude larger than
% FL-NPs in the foam cell. The foam drainage behavior was
a typical surfactant molecule, which implies a large adsorption
observed using a vertical-stage microscope. Figure 11 shows the
energy. The presence of surfactant reduces this energy by about
half because the surface tension decreases from about 70 dyn/
cm to about 30 dyn/cm. This hypothesis suggests that highly
hydrophilic particles do not stabilize interfaces (but they can by
other mechanisms, as discussed below).
Maximum Capillary Pressure of Coalescence. In the
absence of nanoparticles, the foam films are flat, and the
capillary pressure (Pc) is balanced by the disjoining pressure in
the foam film. When the capillary pressure exceeds a threshold
pressure (Pcmax), the film ruptures.51 In the presence of
Figure 11. Images of vertical foam films made using FL-NPs captured nanoparticles, the films do not have to be flat. Nanoparticles
in (A) visible light and (B) UV light (scale bars are 1000 μm). have the potential to provide a steric barrier in the thinning of
foam films and thus play a major role in retarding the
coalescence of foam bubbles. Kaptay52 has derived an
foam structure after 30 min of drainage in visible and UV light. expression for the maximum capillary pressure by analyzing a
From the front view of the vertical films, the cross section of single hexagonal layer of particles between two bubble films:
the horizontal Plateau border, which is perpendicular to the
foam cell, can be seen. The locations of the nanoparticles (i.e., 2γaw
fluorescence) can be easily seen from the image captured in UV Pcmax = β cos θ
R (6)
light. The fluorescence in the Gibbs−Plateau border and the
lamellas between the bubbles indicate nanoparticle trapping in where β is a theoretical packing parameter, γaw is the air−water
these regions. These nanoparticles present a physical barrier in interfacial tension, and R is the particle radius. As the radius
the drainage of free liquid and retard the coalescence process. decreases, the maximum capillary pressure the film can
Confocal Laser Scanning Microscopy. To confirm the experience without rupturing increases. Figure 13 shows a
observations from the vertical foam film tests, CLSM was done spherical particle that bridges two bubble films with contact
473 DOI: 10.1021/ef5015007
Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

Figure 12. Confocal microscopy images of foams stabilized by 0.5 wt % surfactant and 0.3 wt % FL-NPs: (A, D) fluorescence images showing
nanoparticles (green); (B, E) bright-field images; (C, F) merged images (scale bars are 200 μm).

Figure 13. Bridging behaviors of (a) hydrophobic and (b) hydrophilic particles in a foam (adapted from ref 54).

angle θ. When this angle is greater than 90°, the positive using FL-NPs suggest that some of the nanoparticles drain into
capillary pressure in the film adjacent to the particle causes the Plateau borders from foam films. The particles (surfactant
drainage of liquid away from the particle, resulting in dewetting molecules or nanoparticles) are forced to attain an orderly
of the film and the formation of a hole (Figure 13a). However, arrangement within a thinning interbubble film, resulting in
for θ < 90°, as in the present case of hydrophilic nanoparticles, stepwise thinning of the film. This stratification process slows
after an initial drainage a critical film thickness is achieved, and the liquid drainage and provides additional foam stability in
the film becomes planar. Further drainage causes the capillary conjunction with the disjoining pressure. The driving force for
pressure to draw liquid toward the particle, thus stabilizing the the particle to exhibit stepwise thinning is the chemical
film by a bridging mechanism (Figure 13b).53 This theory potential gradient of particle at the film periphery, where the
explains how foam films can be stabilized by bridging particle leaves the interface and a vacancy is formed in its
hydrophilic nanoparticles. place.56 Johonnott57 was the first to show this behavior in foam
Kinetics of Film Drainage. Horozov55 has suggested three film studies using surfactants at high concentrations.
possible ways that hydrophilic particles can be incorporated Sethumadhavan et al.58,59 reported the stratification behavior
inside a liquid film: (a) as a monolayer of bridging particles, (b) of silica nanoparticles resulting in film stability. This tendency
as a bilayer of close-packed particles, and (c) as a network of toward orderly arrangements of particles was reported to
particle aggregates inside the film. Figure 14 illustrates these decrease significantly as the polydispersity of the nanoparticles
three cases in foam films. The images of the vertical foam films increased. The nanoparticles used in the present study were
474 DOI: 10.1021/ef5015007
Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

pressure foam texture as seen in the view cell after the steady
state was reached. The pressure drop has many fluctuations, but
the average pressure drop in the steady state was about 9.7 psi.
The average bubble diameter, measured from the foam texture,
was about 910 μm.
In the subsequent steps, the nanoparticle concentration in
the injection fluid was varied from 0.1 to 0.5 wt % while the
surfactant concentration was kept at 0.5 wt %. The flow rate
was kept at 4 ft/day with 80% quality throughout the
experiments. Steady-state pressure drops were typically
achieved after 8 PV of coinjection. Figure 16 shows the

Figure 14. Three possible mechanisms of liquid film stabilization


(adapted from Horozov55).

highly monodisperse with a very low polydispersity index,


which suggests slow drainage of films.
Foam Flow Experiments. Foam flow experiments were
conducted in 1 ft long Berea sandstone cores using 0.5 wt %
surfactant with various nanoparticle concentrations ranging
from 0.1 to 0.5 wt % as foaming agents. The pressure drop
profiles across different sections of the cores were monitored.
The macroscopic foam texture of effluent coming out of the
core was visualized using the view cell installed at the
downstream end of the core. The flow rates in these
experiments were maintained at 4 ft/day to achieve a pressure Figure 16. Pressure drop profile for the injection of pregenerated foam
drop large enough to be measured accurately. All of these made by coinjecting nitrogen and 0.5 wt % surfactant with 0.1 wt %
experiments were performed at room temperature with a back nanoparticles through a sandpack.
pressure of 100 psi.
First, a surfactant preflush was conducted to saturate the core pressure drop profile when 0.1 wt % nanoparticles was used.
with surfactant and avoid any surfactant adsorption during foam The average steady-state pressure drop was about 13.7 psi,
flooding. The steady-state pressure drop was measured after 3 which is slightly more than in the case with no nanoparticles,
PV of injection and found to be very low (about 0.1 psi). After indicating a stronger foam. Then the nanoparticle concen-
the surfactant preflush, nitrogen gas and surfactant solution tration in the injection fluid was raised to 0.3 wt %. Figure 17
without nanoparticles were coinjected into the sandpack to shows the pressure drop profile and the steady-state foam
generate foam at 4 ft/day with a quality of 80%. This texture. The steady-state pressure drop in this case was 17.4 psi.
pregenerated foam was injected from the top of the core for at It can be seen that the foam in this case consisted of fine-
least 5 PV to achieve a steady-state pressure drop. Figure 15 textured bubbles. The average bubble diameter was found to be
shows the pressure drop obtained for this case and the high-

Figure 17. Pressure drop profile for the injection of pregenerated foam
Figure 15. Pressure drop profile for the injection of pregenerated foam made by coinjecting nitrogen and 0.5 wt % surfactant with 0.3 wt %
made by coinjecting nitrogen and 0.5 wt % surfactant through a nanoparticles through a sandpack. The inset shows the steady-state
sandpack. The inset shows the steady-state foam texture. foam texture.

475 DOI: 10.1021/ef5015007


Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

330 μm, which is smaller than those for the NP-free surfactant 19 shows a plot of effluent history, with the x axis showing the
case. The large pressure drop achieved in this case is due to the PV injected and the y axis showing the dimensionless
finer in situ bubble texture stabilized by the surfactant−
nanoparticle system. With 0.5 wt % concentration of NPs, the
steady-state pressure drop increased to 21.5 psi.
The mobility reduction factor (MRF) is defined here as the
ratio of the pressure drop across the core due to foam flow to
the pressure drop due to single-phase gas flow at the same flow
rate. Figure 18 shows plots of the steady-state pressure drop

Figure 19. Effluent history of nanoparticles.

concentration CD, which is the nanoparticle concentration of


Figure 18. Steady-state pressure drops (black, left axis) and mobility
the effluent normalized by the nanoparticle concentration of
reduction factors (blue, right axis) achieved using various concen- the injected fluid. RNP was calculated using this plot and found
trations of nanoparticles. to be 99.57%. This result along with the data available in the
literature shows that the nanoparticles used in this study have
very low retention in the porous medium and the potential for
and MRF as functions of the nanoparticle concentration. As the long-distance propagation without causing permeability dam-
nanoparticle concentration increased from 0 to 0.5 wt %, the age.
MRF increased from 4000 to 8700. At the end of the Oil Displacement Experiments. Core flood 1 was
experiment, the core was cleaned as described in Methodology, conducted in a Berea core with crude oil C. The initial oil
and the brine permeability was measured. It was found to be saturation was 64%. Figure 20 shows the injection scheme,
335 mD, which is close to the initial value of 357 mD, cumulative oil recovery, and overall pressure drop across the
suggesting minimal permeability damage. A small reduction in core. The brine flood was conducted at 1 ft/day to mimic the
permeability was expected because of some trapped gas in the water flood at a typical field rate. It was continued for 2 PV
core that could not be removed by brine flushing and rapid until no oil was produced. The water flood oil recovery was
pressurization−depressurization cycles. 54.4% original oil in place (% OOIP), and the oil saturation was
Nanoparticle Transport and Retention in Porous reduced to 29.2%. The pressure drop during the water flood
Media. In order to quantify the nanoparticle retention, we was between 1 and 3 psi. Then brine was injected at 5 ft/day
calculated the nanoparticle recovery (RNP), which has been for another 1 PV to minimize the capillary end, if any. The
used in the recent literature.60,61 It is defined as the ratio of pressure drop increased to 7.4 psi as the flow rate was increased
amount of nanoparticles recovered to the amount of nano- 5 times. No oil was recovered during this stage, implying that
particles injected in the system. The quantity of nanoparticles there were no significant capillary end effects. Before the foam
recovered was calculated by integrating the area under the flood was conducted, the core was preflushed with 1 PV of 0.5
effluent history using the trapezoidal approximation. The ratio wt % surfactant solution at 1 ft/day to avoid any surfactant
of this quantity to the number of pore volumes of nanoparticles adsorption during foam flooding. No oil was recovered during
injected multiplied by the nanoparticle concentration in the surfactant injection, as the surfactant did not lower the IFT
injection fluid gives the RNP value. Murphy61 reported retention sufficiently (IFT ≈ 2.3 mN/m) to mobilize residual oil. The
data for Nyacol (DP 9711) nanoparticles, which were used in pressure drop during the preflush was almost constant at 1.1
the present study, in Boise sandstone. He reported RNP values psi. Then coinjection of 0.5 wt % surfactant solution and
of 96% and 95% for injected nanoparticle concentrations of nitrogen gas was started with a quality of 80% at 1 ft/day. Since
2.84 and 1.5 wt %, respectively. In the present study, we there was some dead volume before the core as explained
measured the RNP for a mixture of 0.5 wt % surfactant and 0.5 earlier, the pressure drop increase due to in situ foam
wt % nanoparticles in Berea sandstone under conditions similar generation was delayed by about 0.7 PV. The additional oil
to those used in above foam flow experiments. First, 4 PV of recovery for the first 3.5 PV of coinjection over the water flood
the above mixture (in 1 wt % NaCl brine) was injected in the was 9.3% OOIP. No significant amount of oil was produced
Berea core (previously saturated with brine) at 4 ft/day. after 3.5 PV of injection, but the coinjection was continued for
Second, 4 PV of brine was then injected at same flow rate to another 4.5 PV to observe the foam mobility in the presence of
flush the system. Effluent samples were taken at each 0.1 PV at residual oil. The average pressure drop continued to grow and
each step. These samples were analyzed using ICP-OES. Figure reached about 9 psi at the end of experiment. The ultimate
476 DOI: 10.1021/ef5015007
Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

Figure 20. Pressure drop profile (black, left axis) and cumulative oil recovery (blue, right axis) for core flood 1.

Figure 21. Pressure drop profile (black, left axis) and cumulative oil recovery (blue, right axis) for core flood 2.

cumulative oil recovery was 63.7% OOIP, and the final oil injection was 10.6% OOIP. The coinjection was continued for
saturation was 23.2%. another 4.5 PV. The pressure drop in this core flood went to
Core flood 2 was conducted in another Berea core using the around 9.8 psi at the end of the experiment.
same procedure as for the previous core flood except with both The objective of these oil displacement experiments in water-
surfactant and nanoparticles. The initial oil saturation in this wet cores was to investigate the potential of displacing water
case was 63.8%. The cumulative oil recovery (% OOIP) and flood residual oil by immiscible foams stabilized by surfactant or
overall pressure drop are shown in Figure 21. The core was surfactant−NP blends. These experiments showed that these
flooded with 2 PV of brine at 1 ft/day, which reduced the oil foams can reduce the residual oil by about 10% OOIP. This is
saturation to 29.8% and resulted in 53.3% OOIP oil recovery. significant considering that the gas is immiscible with the oil. In
The pressure drop during this stage was low (∼2 psi). Then the
water-wet systems, water as the most wetting fluid displaces oil
core was flooded with 1 PV of brine at 5 ft/day. No additional
from the small pores and traps oil in the bigger pores. Gas as
oil was recovered, implying that capillary end effects were
negligible. The pressure drop increased to about 10.5 psi at a the most nonwetting fluid tends to occupy the bigger pores and
flow rate of 5 ft/day. Then the core was preflushed with 1 PV of displaces some of that oil. Immiscible foams can increase the
an aqueous solution containing 0.5 wt % surfactant and 0.3 wt trapped gas saturation, which decreases the residual oil
% nanoparticles. No oil was recovered during this stage because saturation, resulting in additional oil recovery. Moreover, they
the oil−water interfacial tension was not ultralow. Coinjection can reduce the gas mobility significantly. It should be noted that
of 0.3 wt % nanoparticles in 0.5 wt % surfactant solution and the cores used in the core floods were only 1 inch in diameter
nitrogen gas was then started at 1 ft/day with 80% quality. The and quite homogeneous; thus, there was no scope to improve
additional oil recovery over the water flood after 3.5 PV of the volumetric sweep efficiency. Similar oil displacement
477 DOI: 10.1021/ef5015007
Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

experiments in heterogeneous and three-dimensional media (3) Orr, F. M., Jr.; Heller, J. P.; Taber, J. J. Carbon dioxide flooding
could address this issue. for enhanced oil recovery: Promise and problems. J. Am. Oil Chem. Soc.


1982, 59 (10), 810A−817A.
CONCLUSIONS (4) Lake, L. W.; Schmidt, R. L.; Venuto, P. B. A niche for enhanced
oil recovery in the 1990s. Oil Gas J. 1990, 88 (17), 62−67.
In this study, the effect of nanoparticle−surfactant blends on (5) Koval, E. J. A method for predicting the performance of unstable
foam performance in both bulk and porous media was miscible displacement in heterogeneous media. Old SPE J. 1963, 3 (2),
evaluated systematically. The following conclusions can be 145−154.
drawn from this work: (6) Kovscek, A. R.; Patzek, T. W.; Radke, C. J. Mechanistic prediction
• Static foam tests indicate a stabilization effect of of foam displacement in multidimensions: A population balance
nanoparticles on surfactant−nanoparticle foam stability approach. SPE/DOE Improved Oil Recovery Symp. 1994,
in the absence of crude oil. An increase in the half-life of DOI: 10.2118/27789-MS.
(7) Rossen, W.; van Duijn, C.; Nguyen, Q.; Shen, C.; Vikingstad, A.
foam with increasing nanoparticle concentration was
Injection strategies to overcome gravity segregation in simultaneous
observed. The foam height remained almost constant for gas and water injection into homogeneous reservoirs. SPE J. 2010, 15
>4 days when the nanoparticle concentration was more (1), 76−90.
than 0.3 wt %. (8) Hirasaki, G. J.; Lawson, J. B. Mechanisms of foam flow in porous
• Vertical foam film tests and confocal laser scanning media: apparent viscosity in smooth capillaries. Old SPE J. 1985, 25
microscopy elucidated that nanoparticles are trapped in (2), 176−190.
the Plateau border as well as lamellas, which retards (9) Bond, D. C.; Holbrook, O. C. Gas drive oil recovery process. U.S.
liquid drainage and bubble coalescence. Patent 2,866,507, 1958.
• The classification of foams on the basis of lamella (10) Patzek, T. Field applications of steam foam for mobility
number was found to be in agreement with the improvement and profile control. SPE Reservoir Eng. 1996, 11 (2), 79−
macroscopic foam−oil interaction behavior. The spread- 86.
(11) Bikerman, J. J. Foams; Springer: New York, 1973; pp 65−97.
ing, entering, and bridging coefficients were found to be
(12) Hirasaki, G. J. Supplement to SPE 19505: The Steam-Foam
unreliable parameters to estimate foam−oil stability. ProcessReview of Steam-Foam Process Mechanisms, 1989.
• As the concentration of nanoparticles increased, the (13) Rossen, W. R. Foams in enhanced oil recovery. Surfactant Sci.
mobility reduction factor of the surfactant−nanoparticle Ser. 1996, 413−464.
foam in a Berea core increased by up to a factor of 2. (14) Grigg, R. B.; Mikhalin, A. A. Effects of Flow Conditions and
• Core floods in a sandstone core with a reservoir crude oil Surfactant Availability on Adsorption. Int. Symp. Oilfield Chem. 2007,
showed that immiscible foams using surfactant or DOI: 10.2118/106205-MS.
surfactant−nanoparticle blends can increase the oil (15) Paul, K. T.; Satpathy, S. K.; Manna, I.; Chakraborty, K. K.;
recovery over water flood by about 10% of the original Nando, G. B. Preparation and characterization of nano structured
oil in place. materials from fly ash: A waste from thermal power stations, by high
energy ball milling. Nanoscale Res. Lett. 2007, 2 (8), 397−404.
These conclusions indicate that nanoparticles can potentially be (16) Alargova, R. G.; Warhadpande, D. S.; Paunov, V. N.; Velev, O.
used to boost the foam performance of surfactant-stabilized D. Foam superstabilization by polymer microrods. Langmuir 2004, 20
foams. Such foams may be able to provide mobility and (24), 10371−10374.
conformance control in miscible and immiscible gas displace- (17) Binks, B. P.; Horozov, T. S. Aqueous foams stabilized solely by
ments. Moreover, this synergy between nanoparticle and silica nanoparticles. Angew. Chem. 2005, 117 (24), 3788−3791.
surfactant in foam stabilization can potentially be used to (18) Gonzenbach, U. T.; Studart, A. R.; Tervoort, E.; Gauckler, L. J.
minimize the surfactant usage and maximize the foam Ultrastable Particle-Stabilized Foams. Angew. Chem., Int. Ed. 2006, 45
propagation distance in subsurface applications. (21), 3526−3530.


(19) Espinoza, D. A.; Caldelas, F. M.; Johnston, K. P.; Bryant, S. L.;
AUTHOR INFORMATION Huh, C. Nanoparticle-stabilized supercritical CO2 foams for potential
mobility control applications. SPE Improved Oil Recovery Symp. 2010,
Corresponding Author DOI: 10.2118/129925-MS.
*Telephone: 512-471-3077. Fax: 512-471-9605. E-mail: (20) Yu, J.; An, C.; Mo, D.; Liu, N.; Lee, R. L. Foam Mobility
[email protected]. Control for Nanoparticle-Stabilized Supercritical CO2 Foam. SPE
Notes Improved Oil Recovery Symp. 2012, DOI: 10.2118/153336-MS.
The authors declare no competing financial interest. (21) Mo, D.; Yu, J.; Liu, N.; Lee, R. L. Study of the Effect of Different


Factors on Nanoparticle-Stabilized CO2 Foam for Mobility Control.
ACKNOWLEDGMENTS SPE Annu. Tech. Conf. Exhib. 2012, DOI: 10.2118/159282-MS.
(22) Aroonsri, A.; Worthen, A. J.; Hariz, T.; Huh, C.; Johnston, K. P.;
We are thankful to the sponsors of the Gas EOR Industrial Bryant, S. L. Conditions for Generating Nanoparticle-Stabilized CO2
Affiliates Project at The University of Texas at Austin and Foams in Fracture and Matrix Flow. SPE Annu. Tech. Conf. Exhib.
Statoil for partial funding of this work. We also thank Dr. Eric 2013, DOI: 10.2118/166319-MS.
Dao for helping us with the experimental setup.


(23) Worthen, A.; Bagaria, H.; Chen, Y.; Bryant, S. L.; Huh, C.;
Johnston, K. P. Nanoparticle Stabilized Carbon Dioxide in Water
REFERENCES Foams for Enhanced Oil Recovery. SPE Improved Oil Recovery Symp.
(1) Taber, J. J.; Martin, F. D.; Seright, R. S. EOR screening criteria 2012, DOI: 10.2118/154285-MS.
revisited-Part 1: Introduction to screening criteria and enhanced (24) Zhang, R.; Somasundaran, P. Advances in adsorption of
recovery field projects. SPE Reservoir Eng. 1997, 12 (3), 189−198. surfactants and their mixtures at solid/solution interfaces. Adv. Colloid
(2) Enick, R. M.; Olsen, D.; Ammer, J.; Schuller, W. Mobility and Interface Sci. 2006, 123−126, 213−229.
Conformance Control for CO2 EOR via Thickeners, Foams, and (25) Zhang, S.; Sun, D.; Dong, X.; Li, C.; Xu, J. Aqueous foams
GelsA Literature Review of 40 Years of Research and Pilot Tests. stabilized with particles and nonionic surfactants. Colloids Surf., A
SPE Improved Oil Recovery Symp. 2012, DOI: 10.2118/154122-MS. 2008, 324 (1−3), 1−8.

478 DOI: 10.1021/ef5015007


Energy Fuels 2015, 29, 467−479
Energy & Fuels Article

(26) Liu, Q.; Zhang, S.; Sun, D.; Xu, J. Foams stabilized by Laponite (48) Li, R. F.; Hirasaki, G.; Miller, C. A.; Masalmeh, S. K. Wettability
nanoparticles and alkylammonium bromides with different alkyl chain Alteration and Foam Mobility Control in a Layered 2D Heteroge-
lengths. Colloids Surf., A 2010, 355 (1−3), 151−157. neous Sandpack. SPE J. 2012, 17 (4), 1−207.
(27) Cui, Z. G.; Cui, Y. Z.; Cui, C. F.; Chen, Z.; Binks, B. P. Aqueous (49) Binks, B. P.; Lumsdon, S. O. Influence of particle wettability on
foams stabilized by in situ surface activation of CaCO3 nanoparticles the type and stability of surfactant-free emulsions. Langmuir 2000, 16
via adsorption of anionic surfactant. Langmuir 2010, 26 (15), 12567− (23), 8622−8631.
12574. (50) Binks, B. P. Particles as surfactantsSimilarities and differences.
(28) Worthen, A. J.; Bryant, S. L.; Huh, C.; Johnston, K. P. Carbon Curr. Opin. Colloid Interface Sci. 2002, 7 (1), 21−41.
dioxide-in-water foams stabilized with nanoparticles and surfactant (51) Denkov, N. D.; Ivanov, I. B.; Kralchevsky, P. A.; Wasan, D. T. A
acting in synergy. AIChE J. 2013, 59 (9), 3490−3501. possible mechanism of stabilization of emulsions by solid particles. J.
(29) Cumberland, S. A.; Lead, J. R. Particle size distributions of silver Colloid Interface Sci. 1992, 150 (2), 589−593.
(52) Kaptay, G. On the equation of the maximum capillary pressure
nanoparticles at environmentally relevant conditions. J. Chromatogr., A
induced by solid particles to stabilize emulsions and foams and on the
2009, 1216 (52), 9099−9105.
emulsion stability diagrams. Colloids Surf., A 2006, 282−283, 387−401.
(30) Diegoli, S.; Manciulea, A. L.; Begum, S.; Jones, I. P.; Lead, J. R.;
(53) Pugh, R. J. Foaming, foam films, antifoaming and defoaming.
Preece, J. A. Interaction between manufactured gold nanoparticles and Adv. Colloid Interface Sci. 1996, 64, 67−142.
naturally occurring organic macromolecules. Sci. Total Environ. 2008, (54) Aveyard, R.; Binks, B. P.; Fletcher, P. D. I.; Peck, T. G.;
402 (1), 51−61. Rutherford, C. E. Aspects of aqueous foam stability in the presence of
(31) Vikingstad, A. K.; Skauge, A.; Høiland, H.; Aarra, M. Foam−oil hydrocarbon oils and solid particles. Adv. Colloid Interface Sci. 1994,
interactions analyzed by static foam tests. Colloids Surf., A 2005, 260 48, 93−120.
(1−3), 189−198. (55) Horozov, T. S. Foams and foam films stabilised by solid
(32) Vikingstad, A. K.; Aarra, M. G.; Skauge, A. Effect of surfactant particles. Curr. Opin. Colloid Interface Sci. 2008, 13 (3), 134−140.
structure on foam−oil interactions: Comparing fluorinated surfactant (56) Kralchevski, P.; Nikolov, A.; Wasan, D. T.; Ivanov, I. Formation
and α-olefin sulfonate in static foam tests. Colloids Surf., A 2006, 279 and expansion of dark spots in stratifying foam films. Langmuir 1990, 6
(1−3), 105−112. (6), 1180−1189.
(33) Zhang, H. (2004). Effect of oils, soap and hardness on the (57) Johonnott, E. S. LXVIII. The black spot in thin liquid films.
stability of foams. Doctoral dissertation, Rice University, Houston, TX, London, Edinburgh Dublin Philos. Mag. J. Sci. 1906, 11 (66), 746−753.
2004. (58) Sethumadhavan, G. N.; Nikolov, A. D.; Wasan, D. T. Stability of
(34) Zou, S.; Yang, Y.; Liu, H.; Wang, C. Synergistic stabilization and liquid films containing monodisperse colloidal particles. J. Colloid
tunable structures of Pickering high internal phase emulsions by Interface Sci. 2001, 240 (1), 105−112.
nanoparticles and surfactants. Colloids Surf., A 2013, 436, 1−9. (59) Sethumadhavan, G.; Nikolov, A.; Wasan, D. Stability of films
(35) Murray, S. B.; Durga, K.; Yusuff, A.; Stoyanov, D. Stabilization with nanoparticles. J. Colloid Interface Sci. 2004, 272 (1), 167−171.
of foams and emulsions by mixtures of surface active food-grade (60) Caldelas, F. M.; Murphy, M.; Huh, C.; Bryant, S. L. Factors
particles and proteins. Food Hydrocolloids 2011, 25 (4), 627−638. governing distance of nanoparticle propagation in porous media. SPE
(36) Peters, E. J. Advanced Petrophysics: Vol. 1: Geology, Porosity, Prod. Oper. Symp. 2011, DOI: 10.2118/142305-MS.
Absolute Permeability, Heterogeneity, and Geostatistics; Live Oak Book (61) Murphy, M. J. Experimental analysis of electrostatic and
Company: Palo Alto, CA, 2012; p 238. hydrodynamic forces affecting nanoparticle retention in porous media.
(37) Simjoo, M.; Nguyen, Q. P.; Zitha, P. L. J. Rheological transition M.S. Thesis, University of Texas at Austin, Austin, TX, 2012.
during foam flow in porous media. Ind. Eng. Chem. Res. 2012, 51 (30),
10225−10231.
(38) Simjoo, M.; Dong, Y.; Andrianov, A.; Talanana, M.; Zitha, P. L.
J. A CT scan study of immiscible foam flow in porous media for EOR.
SPE EOR Conf. Oil Gas West Asia 2012, DOI: 10.2118/155633-MS.
(39) Wang, C.; Bobba, A. D.; Attinti, R.; Shen, C.; Lazouskaya, V.;
Wang, L. P.; Jin, Y. Retention and transport of silica nanoparticles in
saturated porous media: Effect of concentration and particle size.
Environ. Sci. Technol. 2012, 46 (13), 7151−7158.
(40) Andrianov, A.; Farajzadeh, R.; Mahmoodi Nick, M.; Talanana,
M.; Zitha, P. L. J. Immiscible foam for enhancing oil recovery: Bulk
and porous media experiments. Ind. Eng. Chem. Res. 2012, 51 (5),
2214−2226.
(41) Denkov, N. D. Mechanisms of foam destruction by oil-based
antifoams. Langmuir 2004, 20 (22), 9463−9505.
(42) Irani, C. A.; Solomon, C. Slim-Tube Investigation of CO2
Foams. SPE Enhanced Oil Recovery Symp. 1986, DOI: 10.2118/14962-
MS.
(43) Kuhlman, M. Visualizing the effect of light oil on CO2 foams. J.
Pet. Technol. 1990, 42 (7), 902−908.
(44) Jensen, J. A.; Friedmann, F. Physical and Chemical Effects of an
Oil Phase on the Propagation of Foam in Porous Media. SPE
California Regional Meeting 1987, DOI: 10.2118/16375-MS.
(45) Harkins, W. D. A general thermodynamic theory of the
spreading of liquids to form duplex films and of liquids or solids to
form monolayers. J. Chem. Phys. 1941, 9, 552−568.
(46) Robinson, J. V.; Woods, W. W. A method of selecting foam
inhibitors. J. Soc. Chem. Ind. 1948, 67 (9), 361−365.
(47) Schramm, L. L.; Novosad, J. J. Micro-visualization of foam
interactions with a crude oil. Colloids Surf. 1990, 46 (1), 21−43.

479 DOI: 10.1021/ef5015007


Energy Fuels 2015, 29, 467−479

You might also like