Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Colloid and Interface Science 541 (2019) 269–278

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


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

Regular Article

Stabilization of oil-in-water emulsions with graphene oxide and cobalt


oxide nanosheets and preparation of armored polymer particles
Katelynn Edgehouse a, Maria Escamilla a, Louisa Wang b, RhayAuna Dent c, Kevin Pachuta d, Lee Kendall d,
Peiran Wei a, Alp Sehirlioglu d, Emily Pentzer a,⇑
a
Department of Chemistry, Case Western Reserve University, 10900 Euclid Ave, Cleveland, OH 44106, United States
b
Hathaway Brown High School, 19600 N Park Blvd, Shaker Heights, OH 44122, United States
c
John Hay School of Science and Medicine, 2075 Stokes Blvd, Cleveland, OH 44106, United States
d
Department of Materials Science and Engineering, Case Western Reserve University, 10900 Euclid Ave, Cleveland, OH 44106, United States

g r a p h i c a l a b s t r a c t

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

Article history: Pickering emulsions are emulsions stabilized by particles instead of small molecules or polymers, and
Received 27 November 2018 commonly consist of oil droplets dispersed into a continuous water phase with particles lying at the
Revised 13 January 2019 fluid-fluid interface. New particle surfactants are important for tuning the composition and properties
Accepted 22 January 2019
of assemblies and enabling advanced applications, such as energy harvesting and management.
Available online 23 January 2019
Although most particle surfactants are spherical, graphene oxide (GO) nanosheets and clay platelets have
garnered recent attention as 2D (i.e., planar) particle surfactants. Herein, we report the preparation of
Keywords:
Pickering emulsions stabilized by a composite of GO nanosheets and cobalt oxide (CoOx) nanosheets,
Pickering emulsion
Graphene oxide
and illustrate the impact of GO:CoOx ratio, oil identity, and flocculating agent (i.e., salts) on emulsion for-
Cobalt oxide mation and stability. Distinct effects were noted for salt concentration and identity, as well as GO: CoOx
2D particles ratio. We further illustrate the applicability of these GO-CoOx-stabilized emulsions in dispersion polymer-
Dispersion polymerization ization, preparing polystyrene particles armored with both nanosheets. This work provides a method for
facilitating oil-in-water emulsions with composite particle surfactants that are stable for at least a week
and offers the foundation for using the fluid-fluid interface to architect structures of dissimilar materials.
Ó 2019 Elsevier Inc. All rights reserved.

1. Introduction

Emulsions are mixtures of two or more immiscible liquids,


often of substantially different polarity (e.g., oil-water). These
⇑ Corresponding author. systems are of vital importance to a number of daily and
E-mail address: ebp24@case.edu (E. Pentzer). specialized applications, including in food, cosmetics, medicine,

https://doi.org/10.1016/j.jcis.2019.01.092
0021-9797/Ó 2019 Elsevier Inc. All rights reserved.
270 K. Edgehouse et al. / Journal of Colloid and Interface Science 541 (2019) 269–278

etc. [1,2]. Typically, a mixture of two immiscible fluids is agitated, ticle surfactants in Pickering emulsions, but rarely utilized. One
yielding droplets of one dispersed in a continuous phase of the class of TMOs consist of sheets of transition metal atoms and oxy-
other (e.g., oil-in-water emulsions). Most emulsions are kinetically gen that are negatively charged and held together by cations (e.g.,
unstable, and thus upon standing the two phases separate from LiCoO2, LCO) [30]. Such nanosheets of TMOs can be exfoliated from
each other, yielding, for example, an oil phase atop of water [3]. bulk crystals via removal of the cation with strong acid and treat-
Addition of an amphiphilic small molecule such as sodium dodecyl ment with a bulky counterion, for example tetramethylammonium
sulfate (SDS) can render emulsions kinetically stable, although hydroxide (TMAOH). Sehirlioglu and coworkers recently reported
polymers and particles can also be used to lower interfacial tension the chemical exfoliation of cobalt oxide nanosheets (CoOx = sheets
between the two fluids [4]. Of particular interest are Pickering of CoO 2 repeat units bearing some oxygen and cobalt vacancies)
emulsions in which particles reside at the fluid-fluid interface, as from LCO, thoroughly evaluating changes to the nanosheets upon
these systems have stability and applications distinct from emul- exfoliation and giving ready access to 2D CoOx particles [31].
sions stabilized by small molecules or polymers [5]. Further, Pick- Herein, we report the use of GO and CoOx nanosheets as surfac-
ering emulsion platforms have been used to template novel tants for oil-in-water Pickering emulsions, and evaluate the impact
architectures, such as armored polymer particles and both hollow of GO:CoOx ratio, oil identity, flocculating agent identity (inorganic
and filled capsules [6-9]. and organic salt), and concentration. The use of GO and CoOx
A sub-class of Pickering emulsions are those which make use of nanosheets as Pickering emulsion co-surfactants gives access to
2-dimensional (2D) particle surfactants. Provided high aspect ratio novel hybrid structures that can leverage the electrical conductiv-
particles are utilized, the platelets are expected to lie at the fluid- ity of GO (once reduced), and the redox activity of CoOx for appli-
fluid interface, with multiple nanosheets forming a barrier between cations in energy storage and catalysis. For all particle surfactants,
the two fluids, assembling side-by-side and overlapping into multi- nanosheets were flocculated by addition of salt, then oil was added
layers. Common 2D particle surfactants include clay platelets and and the system agitated to form emulsions (Scheme 1). In general,
graphene oxide (GO) nanosheets. Clay platelets are layered struc- the most effective particle surfactant system with the highest load-
tures of aluminum, silicon, and other elements with the faces of ing of CoOx was 1:1 GO:CoOx, and the best flocculating conditions
many clay platelets possessing a negative charge, compensated by utilized LiCl, KCl, or 1-ethyl-3-methylimidazolium hexafluo-
simple cations [10]. Alternatively, GO nanosheets are composed of rophosphate ([Emim][PF6]). Emulsions with different oils were also
a carbon framework bearing alcohol, epoxide, and alkene function- examined. We illustrate that styrene-in-water emulsions stabilized
alities throughout the surface and carboxylic acids on the edges (and by GO-COx can be used to prepare GO-CoOx-armored polymer par-
thus negatively charged on the edges in solution of appropriate pH) ticles by thermally induced free radical dispersion polymerization.
[11]. Clay platelets can assemble at the fluid-fluid interface when Optical microscopy shows individual particles, indicating that GO
flocculated prior to emulsion formation. Upon flocculation with salt, and CoOx are suitable particle surfactants. Further, X-ray photo-
the electrostatic repulsion between neighboring clay platelets is electron spectroscopy (XPS) reveals the presence of cobalt. This
reduced allowing their assembly at different oil-water interfaces work illustrates that GO and CoOx together can stabilize oil-in-
[10,12]. GO nanosheets act as particle surfactants in a similar vein, water emulsions, thereby yielding composites of intimately associ-
however ionic strength, oil identity, pH, GO concentration, and oil: ated nanoparticle structures.
water ratio also impact the system. Specifically, low pH and high
ionic strength yield better emulsion stability [13]. Of note, GO is
2. Materials and methods
chemically heterogeneous as well as reactive [14], and thus the
extent of oxidation and functionalization impact interfacial activity.
2.1. Materials
Indeed, chemical modifications affect the dispersibility of GO and
have been exploited to tailor emulsion identity [15,16]. Recently,
All inorganic salts (>98% purity), oils, and NBu+4Br (1643-19-2,
the Pentzer group modified GO to access octane-in-dimethyl for-
Alfa Aesar, 98%) were purchased from commercial sources
mamide (DMF), DMF-in-octane, ionic liquid (IL)-in-oil, IL-in-water,
(>97% purity) and used as received without further purification.
water-in-IL, and oil-in-IL emulsions [7,8,17,18].
1-ethyl-3-methylimidazolium hexafluorophosphate ([Emim][PF6])
GO nanosheets and clay platelets are the most widely explored
(155371-19-0, 99%), 1-ethyl-3-methylimidazolium dimethyl
2D particle surfactants for Pickering emulsions, leaving a variety of
phosphate ([Emim][DMP]) (945611-27-8, 98.0%), and 1-butyl-
2D particles under-explored. Similar to GO, cellulose nanocrystals
3-methylimidazolium tetrafluoroborate ([Bmim][BF4]) (174501-
(CNCs) are garnering much recent attention for use in Pickering
65-6, 98.0%) were purchased from Iolitec and used as received.
emulsions [19]. Tam and coworkers report chemically modifying
2,20 -Azobis(2-methylpropionitrile) (AIBN, 78-67-1, 98%) was pur-
sulfated CNCs with polystyrene to stabilize toluene-in-water and
chased from Sigma-Aldrich and purified via recrystallization in
hexadecane-in-water Pickering emulsions [20]. More recently,
hot methanol (67-56-1, Fisher Scientific, 99.8%) and dried under
Rojas and coworkers reviewed cellulosic and lignin colloidal parti-
reduced pressure before use. Before dispersion polymerization,
cles at the oil/water interface and their abilities to stabilize Picker-
styrene (100-42-5, Sigma-Aldrich, 98%) was passed through a basic
ing emulsions [21]. Miele and coworkers reported the use of 2D
alumina column to remove inhibitor. Graphite flakes (7782-42-5)
hexagonal boronitride to stabilize water-in-ethyl benzoate Picker-
and sulfuric acid (7664-93-9, 95-98%) were purchased from Sigma
ing emulsions, and used this emulsion platform to template porous
Aldrich, and hydrogen peroxide (7722-84-1, 35 wt% in water) was
membranes [22,23]. Another interesting class of 2D materials is
purchased from Acros Organics. Potassium permanganate (KMnO4,
transition metal dichalcogenides (TMDs), including MoS2, WS2,
7722-64-7, 99.0%) was purchased from Alfa Aesar. Bulk lithium
and MoSe2 [24-26]. TMDs have been widely explored in digital
cobalt oxide (LCO, 12190-9-3) was purchased from MTI Corpora-
electronic and optoelectronic devices [27], but not studied as par-
tion, and tetramethylammonium hydroxide (TMAOH, 75-59-2,
ticle surfactants. More recently, MXenes have emerged as a class of
25 wt% in water) was purchased from Alfa Aesar.
attractive 2D materials; these are accessed by exfoliation of ternary
carbide and nitride ceramics [28]. Huang and coworkers reported
using Ti3C2-MXene to stabilize neutral and basic oil-in-water high 2.2. Instrumentation
internal phase emulsions (HIPEs), and then used these assemblies
to prepare a solid porous monolith [29]. Like TMDs and MXenes, XPS data was collected using a PHI Versaprobe 5000 scanning
transition metal oxides (TMOs) are 2D materials attractive as par- X-ray photoelectron spectrometer. The data was processed using
K. Edgehouse et al. / Journal of Colloid and Interface Science 541 (2019) 269–278 271

Scheme 1. Overview figure illustrating the formation of oil-in-water emulsions stabilized by graphene oxide (GO), cobalt oxide (CoOx), or GO and CoOx (GO/CoOx), as well as
representative photographs and optical images of the resulting emulsions.

Origin 2018 64bit and the counts per second (CPS) normalized H2SO4 (150 mL) and the solution stirred at room temperature
using the Z scores (standardized to N(0.1)). Centrifugation was per- using a magnetic stir bar. Then, an aliquot of KMnO4 (1.0 g,
formed with an Eppendorf 5804 centrifuge. Ultrasonication was 63 mmol) was carefully added to the graphite/H2SO4 mixture and
done with a Branson M3800 bath sonicator, and vortex mixing stirred for 24 h. This procedure was repeated three more times
was achieved with a Fisher Model 9454FIFSUS vortex mixer. until a total of 4 g of KMnO4 had been added. The solution was
Atomic force microscopy (AFM) was performed using a NX-10 Park quenched by transferring to an ice-water solution (0.75 L), fol-
System in tapping mode for GO and contact mode for CO and lowed by the addition of aqueous H2O2 until the color changed
imaged in topography mode. Samples for AFM analysis were pre- from pink to yellow, indicating that excess KMnO4 was quenched.
pared by drop casting from solution onto mica. Fourier transform Crude GO was isolated as a yellow-brown solid via centrifugation.
infrared (FTIR) spectroscopy was performed using an Agilent Cary The supernatant was discarded, and the solid GO washed repeat-
630 FTIR with a diamond/ZnSe crystal in ATR mode. Zeta potential edly with 2-propanol (67-63-0, Fisher Scientific, 99.9%) until the
measurements were performed on a Mobius, Wyatt Technology. supernatant achieved a neutral pH. GO was then dried under
Scanning electron microscopy (SEM) measurements were taken reduced pressure at room temperature. To prepare a stock GO solu-
on a VEGA3 TESCAN instrument. For imaging of GO-CoOx coated tion, solid GO was vortexed and sonicated in water at a concentra-
polystyrene (PS) particles, the voltage was 20.5 kV and for imaging tion of 1 mg/mL.
of GO-CoOx nanosheets 11.0 kV. Optical microscopy images were
taken using an AmScope M150C Microscope with an AmScope
2.4. Preparation of Cobalt Oxide (CoOx)
MD35 camera. Samples for optical microscopy were prepared by
drop casting emulsions on a glass slide and spreading with a pip-
CoOx was synthesized following a previously reported method
ette tip, then application of a top slide. Optical microscopy images
[31]. LCO powder (3.30 g) was dispersed in HCl (7647-01-0, VWR
were collected at the center of the sample. Emulsion stability was
Analytical, 36.5%–38%) (330 mL, 1.0 M) and magnetically stirred
tested by letting the emulsions stand unagitated for a period of
for 24 h at room temperature. The solid was isolated by gravity fil-
time and then taking optical microscopy images.
tration, then washed with water (650 mL) and allowed to air dry
overnight. The dried solid (0.15 g) was then dispersed in aqueous
2.3. Preparation of Graphene Oxide (GO) TMAOH (15 mL of 1.25 vol%). This mixture was then agitated using
a mechanical stir bar at room temperature for 5 days. The solution
GO was synthesized following a method previously reported was then centrifuged at 4000 rpm for 5 min and then the super-
[11]. Briefly, graphite flakes (1.0 g) were dispersed in concentrated natant was separated and again centrifuged for 10 min. The super-
272 K. Edgehouse et al. / Journal of Colloid and Interface Science 541 (2019) 269–278

natant was decanted, and the pellet dried under reduced pressure and a few microns in diameter whereas CoOx nanosheets are
at room temperature. To prepare a stock CoOx solution, the dried 0.8 nm thick and 1 mm in diameter. GO and CoOx both have a
solid was dispersed in water at a concentration of 1 mg/mL negative Zeta potential, showing that both nanosheets had a nega-
(agitated by vortex). tive charge. Both GO and CoOx were dispersed in water at a con-
centration of 1 mg/mL. As a dispersion, GO was more stable than
2.5. Preparation of Oil-in-Water Pickering Emulsions Stabilized by GO, CoOx upon standing, as the latter precipitated after 5 h.
CoOx, and GO- CoOx First, the ability to form toluene-in-water Pickering emulsions
with only GO, only CoOx, or different ratios of the nanosheets as
Standard oil-in-water emulsions were prepared by dispersing particle surfactant was evaluated using a 2:3 toluene:water ratio
GO in distilled water (1.0 mg/mL) and CoOx in distilled water (a similar system has been reported for GO). As expected, GO
(1.0 mg/mL). A total of 4 mL of aqueous nanosheet solution was nanosheets formed emulsions (Fig. 1A), however CoOx nanosheets
placed in a 2 dram vial with the following GO:CoOx ratios – 1:0, did not, and simply aggregated (Fig. 1F). Further studies revealed
2:1, 1:1, 1:2, 1:5, and 0:1. This solution (0.75 mL) was then placed the GO:CoOx ratio dictated the ability to form an emulsion, with
in a 2 mL vial. Oil (0.50 mL) was added to the vial, and then the vial 2:1, 1:1, and 1:2 GO:CoOx ratios leading to the formation of dro-
was vortexed for 5 s and bath sonicated for 30 s [8]. All emul- plets (Fig. 1B–D), but a higher content of CoOx leading again to
sions were prepared in triplicate. aggregation of nanosheets (1:5 GO:CoOx, Fig. 1E). Although the
presence of enough GO led to emulsion formation for GO:CoOx sys-
tems, nanosheets remained in the continuous phase, emulsion dro-
2.6. Preparation of Oil-in-Water Pickering Emulsions Stabilized by GO,
plets varied in size, and the emulsions were dramatically less
CoOx, and GO and CoOx with Salt
stable than emulsions stabilized by GO only. The inability of CoOx
and 1:5 GO:CoOx to stabilize toluene-in-water emulsions suggests
A given amount of salt was weighed into a 2 dram vial, then the
that CoOx nanosheets do not have a suitable affinity for the oil-
nanosheet solution (4.0 mL) was added to the vial, and the vial vor-
water interface and cannot act as particle surfactants. Whereas
texed for 10 s. Next, this solution (0.75 mL) was transferred to a
GO and CoOx both have net negative charge (see above), the nega-
2 mL vial, and oil (0.50 mL) was added. The vial was vortexed for
tive charge of GO is attributed to carboxylic acid groups present
5 s and sonicated for 30 s. For CaCl2 (10043-52-4, Merck) and
only at the nanosheet edges, and thus inter-sheet repulsion can
MgCl2 (7786-30-3, Alfa Aesar, 99%), the salts were added to the
be overcome by offset GO nanosheets when assembled at the oil-
nanosheet solutions as a saturated aqueous solution until floccula-
water interface. In contrast, the negative charge of CoOx
tion was observed; the method was altered because dissolution of
nanosheets is likely distributed throughout the faces and edges,
these salts in water is highly exothermic. To evaluate reproducibil-
which renders them more dispersible in water.
ity, all emulsions were prepared in triplicate.
Flocculation of negatively charged nanosheets can facilitate
their assembly at the oil-water interface and use as surfactants.
2.7. Preparation of GO-CoOx Armored Polystyrene Particles via For example, the groups of Bon and Bourgeat-Lami illustrated that
Dispersion Polymerization addition of salt improves the ability of clay platelets (e.g., Laponite)
to stabilize oil-in-water Pickering emulsions [12,33,34]. Addition
GO-CoOx armored PS particles were synthesized via dispersion of NaCl compresses the double layer, lowering the electrostatic
polymerization of styrene-in-water with a GO:CoOx ratio of 1:1 repulsion between nanosheets, causing flocculation, and leading
[32]. LiCl (0.15 g), GO stock solution (3 mL), and CoOx stock solu- to the formation of multiple layers of clay at the oil-water interface
tion (3 mL) were added to a 25 mL round bottom flask,. The flask [10]. Thus, in an effort to facilitate the formation of Pickering emul-
was swirled until the salt dissolved. Then, AIBN (0.3 g, 1.8 mmol) sions stabilized by GO-CoOx and CoOx, the nanosheets were floccu-
was dispersed in styrene (3 mL), and then this solution was added lated with readily available inorganic and organic salts. Salt was
to the round bottom flask. The mixture was sonicated for 30 min in added to an aqueous suspension of nanosheets to induce floccula-
cold water, and then degassed by bubbling nitrogen through the tion, then toluene was added and the resulting mixture agitated by
solution. The resulting emulsion was placed under nitrogen gas vortex and sonication to yield opaque, milky emulsions. This
and heated to 70 °C without stirring. After 24 h, the reaction was approach dramatically facilitated the formation of emulsions. The
cooled to room temperature to yield a milky gray solution. The quality of emulsions was determined by droplet uniformity (spher-
styrene:water ratios used were 1:2, 1:3, and 1:5, with 1:3 yielding ical shape and diameter heterogeneity) and lack of nanosheets in
discrete particles. the continuous phase, whereas stability was tested by letting the
emulsions to stand unagitated for 10 weeks and comparing pho-
3. Results and discussion tographs and optical microscopy images.
The inorganic salts used to flocculate GO, GO-CoOx, and CoOx
Graphene oxide (GO) and cobalt oxide (CoOx) nanosheets were prior to emulsion formation were LiCl, NaCl, KCl, CsCl, MgCl2, and
prepared as previously reported, and thoroughly characterized by a CaCl2; a variety of salt concentrations were evaluated, as well as
number of techniques (see Fig. S1, S2, and Table S1). Briefly, GO different GO:CoOx ratios. Figs. S3–S8 shows the impact of salt con-
nanosheets were prepared by the chemical oxidation of expanded centration and GO:CoOx ratio for each inorganic salt evaluated;
graphite using a modified Hummer’s method [11]. The FTIR Fig. 2 shows photographs and optical microscopy images of the
spectrum of GO nanosheets revealed the characteristic stretching best emulsions for each salt. For each system, 1:1 GO:CoOx yielded
frequencies of GO associated with OAH, C@C, and C@O bonds; the best emulsions, though the optimum concentration was differ-
X-ray photoelectron spectroscopy (XPS) revealed an overall C:O ent for each salt. The addition of monovalent chloride salts (LiCl,
ratio of 2:1 and Raman spectroscopy showed the signature D and NaCl, KCl, CsCl, Fig. 2A–D) all improved emulsion formation, as
G bands of GO at 1350 cm1 and 1600 cm1 respectively (2:1, indicated by more homogenous droplet diameters and a clear con-
AD/AG). The CoOx nanosheets were prepared by the chemical exfo- tinuous phase, indicating no nanosheets remained in the water and
liation of commercially available LCO by treatment with aqueous suggesting all are associated with the oil-water interface. In con-
HCl then TMAOH. XPS was used to monitor this exfoliation process, trast, neither the addition of MgCl2 nor CaCl2 led to improved
specifically by following the Co 2p peak at 780 eV. Atomic force emulsion formation (heterogeneous droplet sizes, compare
microscopy (AFM) images reveal GO nanosheets are 2 nm thick Fig. 1C, Fig. 2E and F).
K. Edgehouse et al. / Journal of Colloid and Interface Science 541 (2019) 269–278 273

Fig. 1. Optical microscopy images and photographs of toluene-in-water emulsions stabilized by (A) GO, (B) 2:1 GO:CoOx, (C) 1:1 GO:CoOx, (D) 1:2 GO:CoOx, (E) 1:5 GO:CoOx,
and (F) CoOx.

Fig. 2. Optical microscopy images and photographs of toluene-in-water emulsions prepared with GO:CoOx (1:1) flocculated with different inorganic salts: (A) 0.59 M LiCl; (B)
0.43 M NaCl; (C) 0.17 M KCl; (D) 0.74 M CsCl; (E) 0.06 M MgCl2; and (F) 0.34 M CaCl2.

Across the inorganic salts evaluated, all facilitated the forma- water emulsions. Similar trends were observed for other inorganic
tion of emulsion droplets with all ratios of GO:CoOx, including salts (Figs. S4–S8).
CoOx-only and 1:5 GO:CoOx. These data support that the salt helps XPS was used to confirm the presence of GO and CoOx in these
shield the negative charges of neighboring nanosheets and facili- composite emulsion surfactants. An emulsion with 1:1 GO:CoOx
tates their assembly at the oil-water interface. Fig. 3 illustrates flocculated with LiCl was prepared and then the composite
the impact of the addition of LiCl on emulsion formation for differ- nanosheet mixture isolated by centrifugation, with the clear super-
ent nanosheet systems. In contrast to the data discussed above in natant discarded. Fig. 4A–C compare the survey and high-
which no salt was added, for GO, CoOx, and 2:1, 1:1, 1:2, and 1:5 resolution C 1s and O 1s XPS spectra of LCO (blue spectra), GO
GO:CoOx, droplets are observed. At lower GO:CoOx ratios (1:2 (red spectra), and GO:CoOx (black spectra). The survey spectrum
and 1:5, Fig. 3D and E), nanosheets are apparent in the continuous of GO:CoOx reveals the presence of oxygen (pink box), carbon
phase, and with the CoOx-only system nanosheet aggregates dom- (orange box), and chloride salt (blue box, Cl 2p), as well as a small
inate, although some droplets are present (Fig. 3F). Neither lower signal due to cobalt (purple box). The high-resolution C 1s spectra
nor higher salt concentration led to all CoOx nanosheets associated shown in Fig. 4B support the presence of GO in the GO-CoOx pellet,
with the oil-water interface, and the highest salt concentration led with a C:O ratio of 1:1. Fig. 4C shows the high-resolution Co 2p
to nanosheet precipitation (see Fig. S3). These observations suggest spectra, illustrating the presence of cobalt in the GO:CoOx compos-
that CoOx-only flocculates are not suitable for stabilizing oil-in- ite (peaks at 780 eV and 795 eV). Characterization of the GO:CoOx
274 K. Edgehouse et al. / Journal of Colloid and Interface Science 541 (2019) 269–278

Fig. 3. Optical microscopy images and photographs of toluene-in-water emulsions prepared with different GO-CoOx composition, flocculated with 0.59 M LiCl.

Fig. 4. Characterization of LCO (blue traces), GO (red traces), and GO-CoOx composites (1:1 ratio, black traces): (A) Survey XPS spectra; (B) High resolution C 1 s spectra; (C)
High resolution Co 2p spectra; (D) SEM of the GO-CoOx composites. (For interpretation of the references to color in this figure legend, the reader is referred to the web version
of this article.)

composite by SEM shows a rough morphology, and possible layer- plimentary system [35]. The flocculated nanosheets and emulsions
ing of the two types of nanosheets (Fig. 4D). were prepared in a similar manner to those discussed above
In addition to evaluating inorganic salts as flocculating agents (1 mg/mL solutions of GO and CoOx, and GO:CoOx ratios of 2:1,
for GO and CoOx nanosheets, water-miscible organic salts were 1:1, 1:2, 1:5, GO-only, and CoOx-only, Fig. S9–S12). [Emim][PF6],
examined. In comparison to their inorganic counterparts, the 1-ethyl-3-methylimidazolium dimethylphosphate ([Emim]
charge on organic salts is more delocalized, and thus the molecule [DMP]), 1-butyl-3-methylimidazolium tetrafluoroborate ([Bmim]
softer and more polarizable; furthermore, organic salts are less [BF4]), and tetra-n-butylammonium bromide (NBu+4Br) were all
hydrophilic than their inorganic counterparts and provide a com- evaluated as flocculating agents. Addition of [Emim][PF6] and
K. Edgehouse et al. / Journal of Colloid and Interface Science 541 (2019) 269–278 275

[Emim][DMP] led to the formation of emulsions with relatively After evaluating the ability of various inorganic and organic
homogenous droplets (Fig. 5A and B), NBu+4Br led to the formation salts to facilitate the formation of emulsions with GO and CoOx
of an emulsion with a large distribution of droplet sizes (Fig. 5C), nanosheets as surfactant, stability of the toluene-in-water systems
and addition of [Bmim][BF4] did not lead to emulsion formation, was studied by leaving the emulsions unagitated. From the discus-
only nanosheet aggregation (Fig. S9). In comparison to [Emim] sion above, flocculation of 1:1 GO:CoOx with LiCl, KCl, or [Emim]
[PF6] and [Emim][DMP], the cations of [Bmim][BF4] and NBu+4Br [PF6] led to the formation of the best emulsions and thus these sys-
are more hydrophobic, having longer alkyl chains (butyl vs. ethyl) tems were studied. Fig. 6 shows the optical microscopy images and
which also render them bulkier; these differences may account for photographs of the emulsions as prepared, after 1 day, after
the lack of emulsion formation. As in the case of inorganic salts, 1 week, and after 10 weeks. After 1 week unagitated, the droplets
nanosheet ratio impacts emulsion formation: emulsions with more of all samples appeared to increase slightly in diameter, but
GO tend to have little-to-no nanosheets remaining in the continu- remained spherical, with the LiCl droplets undergoing the most
ous phase, and emulsions with lower GO:CoOx ratios (1:2, 1:5, and dramatic change. The difference in droplet size and size distribu-
CoOx-only) had larger and more disperse droplet sizes as well as tion appeared similar for all samples between 1 day and 1 week.
nanosheets in the continuous phase. In contrast, after 10 weeks of standing unagitated, the KCl emul-
sion lost integrity, with some nanosheets precipitating from solu-
tion and the [Emim][PF6] emulsion underwent catastrophic
changes, with little-to-no emulsion droplets remaining and a large
amount of precipitation. In contrast, after 10 weeks no changes
were observed for the LiCl emulsion. The emulsion stability
indexes of the LiCl and KCl-flocculated emulsions were determined
by the volumetric method to be 7.1% and 18.8%, respectively,
whereas that of the [Emim][PF6] emulsions was 0% (see Fig. S17)
[36]. These samples indicate that the salt used to flocculate the
nanosheets directly impacts the stability of the resulting emulsion,
as each contained the same amount and ratio of GO:CoOx and
toluene-to-water.
Once the preparation of toluene-in-water emulsions stabilized
by GO-CoOx was established, other oil phases were tested: hex-
anes, dodecane, chloroform, and styrene. A variety of salts were
evaluated for each oil, and the best emulsions formed shown in
Fig. 7. Hexanes-in-water emulsions were remarkably different
from those with dodecane, notable given that both oils are hydro-
carbons and thus of similar polarity. Emulsion droplets for
Fig. 5. Optical microscopy images and photographs of toluene-in-water emulsions
hexanes-in-water were larger with a broader size distribution than
prepared when aqueous solutions of GO:CoOx (1:1) are flocculated with different dodecane-in-water emulsion droplets (compare Fig. 7A and B);
organic salts: (A) 0.10 M [Emim][DMP]; (B) 0.20 M [Emim][PF6]; C) 0.20 M NBu+4Br. these differences could be attributed to boiling points of the sol-
vents (hexanes b.p. = 68 °C and dodecane b.p. = 216 °C) or viscosity

Fig. 6. Stability testing of 1:1 GO:CoOx toluene-in-water emulsions with 0.59 M LiCl, 0.17 M KCl, and 0.20 M [Emim][PF6] as prepared (0 h, 1 day (24 h), 1 week (168 h), and
10 weeks.
276 K. Edgehouse et al. / Journal of Colloid and Interface Science 541 (2019) 269–278

GO-CoOx-stabilized styrene-in-water emulsions are of interest


for the preparation of armored polymer particles, as styrene is
polymerizable and has been applied to dispersion and emulsion
polymerizations [37-39]. For example, Thickett and Zetterlund
reported the miniemulsion polymerization of GO-stabilized
styrene-in-water emulsions to prepare GO-armored polystyrene
(PS) particles [40]. In a similar vein, Bon and coworkers utilized
this technique to fabricate polymer particles armored with
Laponite clay platelets [10,34,41]. Further, Bourgeat-Lami and
coworkers used the Pickering emulsion platform to make armored
co-polymer particles [42]. Here, we used a GO:CoOx ratio of 1:1,
the nanosheets were flocculated with LiCl, and the oil phase con-
sisted of styrene with 0.1 wt% 2.20 -azobis(2-methylpropionitrile)
(AIBN), a thermally activated radical initiator. For these emulsions,
a 1:3 styrene:water ratio was used. After degassing, polymeriza-
tion was initiated by heating the emulsion to 70 °C for 24 h. This
resulted in an opaque, gray solution, and the solid polymer parti-
cles were isolated by filtration, then characterized by XPS, scanning
Fig. 7. Optical microscopy images and photographs of different oil-in-water
emulsions. The salts used for flocculation were (A) 0.20 M [Emim][PF6], (B)
electron microscopy (SEM), and optical microscopy.
0.17 M KCl, (C) 0.59 M LiCl, and (D) 0.17 M KCl. Optical microscopy images revealed individual GO-CoOx-
armored PS particles, with the particle diameters similar to the
styrene-in-water droplets prior to polymerization (Fig. 8A). This
(hexanes = 0.297 cP and dodecane = 1.374 cP). However, it has observation indicates that the GO-CoOx nanosheet composite is a
been shown that the best oils for GO-stabilized Pickering suitable surfactant to contain the inner styrene phase, preventing
emulsions are those that are polar and aromatic [13], which could droplet aggregation or coalescence during heating. The particles
explain the poor performance of hexanes due to its low polarity were further characterized by SEM, after isolation by gravity filtra-
and lack of aromaticity. Chloroform-in-water and styrene-in- tion and affixing to carbon tape. SEM images reveal a textured sur-
water emulsions both readily formed without nanosheets in the face of the GO-CoOx-coated PS particles (Fig. 8B), similar to that
continuous phase, though the droplets were larger and more observed for GO-or clay-armored polymer particles. Characteriza-
heterogeneous in diameter compared to toluene-in-water or tion of the GO-CoOx-coated PS particles by XPS illustrated the pres-
dodecane-in-water emulsions. Evaluation of the impact of GO: ence of cobalt, as expected, and as indicated by peaks associated
CoOx ratio on the formation of chloroform-in-water emulsions with Co 2p (Fig. 8C and Fig. S18, 795 and 780 eV, similar to the
revealed that at increased loading of CoOx, irregular, multi- GO-CoOx mixture itself). As XPS is a surface sensitive technique,
emulsions were formed regardless of salt identity (water-in-oil- the detection of Co confirms the nanosheets lay near the surface
in-water, Fig. S13–S16). of the particle.

Fig. 8. Characterization of GO-CoOx-armored polystyrene particles: (A) (i) optical microscopy images of the styrene-in-water emulsion pre-polymerization, and (ii) post
polymerization; (B) SEM image of GO-CoOx coated PS particle with inset showing single particle; and (C) High resolution Co 2p XPS spectrum.
K. Edgehouse et al. / Journal of Colloid and Interface Science 541 (2019) 269–278 277

4. Conclusion encapsulation of phase change materials in carbon nanoparticles, J. Mater.


Chem. A 6 (2018) 2461–2467.
[10] S. Cauvin, P.J. Colver, S.A.F. Bon, Pickering stabilized miniemulsion
In summary, we have prepared a variety of oil-in-water emul- polymerization: preparation of clay armored latexes, Macromolecules 38
sions stabilized by a combination of graphene oxide (GO) and (19) (2005) 7887–7889.
[11] D.C. Marcano, D.V. Kosynkin, J.M. Berlin, A. Sinitskii, Z.Z. Sun, A. Slesarev, L.B.
cobalt oxide (CoOx) nanosheets. We found that the nanosheets
Alemany, W. Lu, J.M. Tour, Improved synthesis of graphene oxide, ACS Nano 4
must be flocculated with salt prior to emulsion formation and that (8) (2010) 4806–4814.
the best particle surfactant system was 1:1 ratio of GO:CoOx; at [12] B. Brunier, N. Sheibat-Othman, M. Chniguir, Y. Chevalier, E. Bourgeat-Lami,
Investigation of four different laponite clays as stabilizers in pickering
higher CoOx composition, nanosheet aggregates formed, and few
emulsion polymerization, Langmuir 32 (24) (2016) 6046–6057.
emulsion droplets were realized. Various inorganic and organic [13] Y. He, F. Wu, X. Sun, R. Li, Y. Guo, C. Li, L. Zhang, F. Xing, W. Wang, J. Gao,
salts were evaluated as flocculating agents, and LiCl was deter- Factors that affect pickering emulsions stabilized by graphene oxide, ACS Appl.
mined to facilitate emulsion formation, leading to relatively Mater. Interf. 5 (11) (2013) 4843–4855.
[14] A De Leon, M. Mellon, J. Mangadlao, E. Pentzer, The PH dependent reactions of
homogenous droplets, and giving prolonged stability (emulsions graphene oxide with small molecule thiols, RSC Adv. 8 (2018) 18388–18395.
remained after 10 weeks unagitated). Further, a variety of oils were [15] B.J. Rodier, A. De Leon, C. Hemmingsen, E. Pentzer, Polymerizations in oil-in-oil
used, including a monomer that can be polymerized. Dispersion emulsions using 2D nanoparticle surfactants, Polym. Chem. 9 (13) (2018)
1547–1550.
polymerization of the styrene-in-water system led to the forma- [16] Q. Luo, Y. Wang, E. Yoo, P. Wei, E. Pentzer, Ionic liquid-containing pickering
tion of discrete polystyrene particles armored with GO and CoOx, emulsions stabilized by graphene oxide-based surfactants, Langmuir 34 (34)
illustrating the integrity of the Pickering emulsions using com- (2018) 10114–10122.
[17] B. Rodier, A. De Leon, C. Hemmingsen, E. Pentzer, Controlling oil-in-oil
bined particle surfactants. Versatility in oil choice and nanoparticle pickering-type emulsions using 2D materials as surfactant, ACS Macro. Lett. 6
surfactants offers a pathway to template new hybrid architectures (11) (2017) 1201–1206.
including capsules, other armored particles, and HIPEs [43,44]. The [18] B.T. McGrail, J.D. Mangadlao, B.J. Rodier, J. Swisher, R. Advincula, E. Pentzer,
Selective mono-facial modification of graphene oxide nanosheets in
combined use of reduced GO and CoOx will find application in elec-
suspension, Chem. Commun. 52 (2) (2016) 288–291.
trochemical devices (such as for energy storage) given the electri- [19] E Ben Ayed, R. Cochereau, C. Dechance, I. Capron, T. Nicolai, L. Benyahia,
cal conductivity of the former [45,46] and the redox activity of the Water-in-water emulsion gels stabilized by cellulose nanocrystals, Langmuir
34 (2018) 6887–6893.
latter [47,48].
[20] C. Tang, S. Spinney, Z. Shi, J. Tang, B. Peng, J. Luo, K.C. Tam, Amphiphilic
cellulose nanocrystals for enhanced pickering emulsion stabilization,
Langmuir 34 (2018) 12897–12905.
5. Notes
[21] L. Bai, L.G. Greca, W. Xiang, J. Lehtonen, S. Huan, R.W.N. Nugroho, B.L. Tardy, O.
J. Rojas, Adsorption and assembly of cellulosic and lignin colloids at oil/water
The authors declare no competing financial interest. interfaces, Langmuir (2018).
[22] D. Gonzalez Ortiz, C. Pochat-Bohatier, J. Cambedouzou, S. Balme, M. Bechelany,
P. Miele, Inverse pickering emulsion stabilized by exfoliated hexagonal-boron
Acknowledgements nitride (h-BN), Langmuir 33 (46) (2017) 13394–13400.
[23] D. Gonzalez Ortiz, C. Pochat-Bohatier, S. Gassara, J. Cambedouzou, B. Mikhael,
P. Miele, Development of novel H-BNNS/PVA porous membranes via pickering
K.E., M.E., P.W., and E.P. thank NSF CAREER Award no. 1551943 emulsion templating, Green Chem. 20 (2018) 4319–4329.
for financial support. R.D. thanks the ACS SEED program for finan- [24] J. Kim, S. Byun, A.J. Smith, J. Yu, J. Huang, Enhanced electrocatalytic properties
cial support. The work of K.P. and A.S. was supported by the Air of transition-metal dichalcogenides sheets by spontaneous gold nanoparticle
decoration, J. Phys. Chem. Lett. 4 (8) (2013) 1227–1232.
Force Office of Scientific Research (AFOSR FA9550-18-1-0030).
[25] M. Chhowalla, H.S. Shin, G. Eda, L.J. Li, K.P. Loh, H. Zhang, The Chemistry of
The authors also thank the Swagelock Center for Surface Analysis two-dimensional layered transition metal dichalcogenide nanosheets, Nat.
of Materials (SCSAM) for XPS and the CWRU Physics Department Chem. 5 (4) (2013) 263–275.
[26] W.L. Zhang, D. Jiang, X. Wang, B.N. Hao, Y.D. Liu, J. Liu, Growth of polyaniline
for SEM.
nanoneedles on MoS2 nanosheets, tunable electroresponse, and
electromagnetic wave attenuation analysis, J. Phys. Chem. C 121 (9) (2017)
4989–4998.
Appendix A. Supplementary material
[27] N.K. Oh, H.J. Lee, K. Choi, J. Seo, U. Kim, J. Lee, Y. Choi, S. Jung, J.H. Lee, H.S. Shin,
et al., Nafion-mediated liquid-phase exfoliation of transition metal
Supplementary data to this article can be found online at dichalcogenides and direct application in hydrogen evolution reaction,
https://doi.org/10.1016/j.jcis.2019.01.092. Chem. Mater. 30 (14) (2018) 4658–4666.
[28] C. Han, Q. Cui, P. Meng, E.R. Waclawik, H. Yang, J. Xu, Direct observation of
carbon nitride-stabilized pickering emulsions, Langmuir 34 (2018) 10135–
References 10143.
[29] R. Bian, R. Lin, G. Wang, G. Lu, W. Zhi, S. Xiang, T. Wang, P.S. Clegg, D. Cai, W.
Huang, 3D assembly of Ti3C2-MXene directed by water/oil interfaces,
[1] M. Chappat, Some applications of emulsions, Coll. Surf. A Physicochem. Eng.
Nanoscale 10 (8) (2018) 3621–3625.
Asp. 91 (1994) 57–77.
[30] S. Lee, X. Jin, I.Y. Kim, T.H. Gu, J.W. Choi, S. Nahm, S.J. Hwang, Superior additive
[2] X. Wang, M. Zeng, Y.H. Yu, H. Wang, M.S. Mannan, Z. Cheng, Thermosensitive
of exfoliated RuO2 nanosheet for optimizing the electrode performance of
ZrP-PNIPAM pickering emulsifier and the controlled-release behavior, ACS
metal oxide over graphene, J. Phys. Chem. C 120 (22) (2016) 11786–11796.
Appl. Mater. Interf. 9 (8) (2017) 7852–7858.
[31] Pachuta, K. G.; Pentzer, E. B.; Sehirlioglu, A. Compositional Changes Associated
[3] I. Capek, Degradation of kinetically-stable o/w emulsions, Adv. Coll. Interf. Sci.
with the Chemical Exfoliation of Lithium Cobalt Oxide into Atomically Thin
107 (2–3) (2004) 125–155.
CoO2 Nanosheets. In Revisions.
[4] K.P. Velikov, O.D. Velev, K.G. Marinova, G.N. Constantinides, Effect of the
[32] X. Song, Y. Yang, J. Liu, H. Zhao, PS colloidal particles stabilized by graphene
surfactant concentration on the kinetic stability of thin foam and emulsion
oxide, Langmuir 27 (3) (2011) 1186–1191.
films, J. Chem. Soc. Trans. 93 (11) (1997) 2069–2075.
[33] A. Lotierzo, S.A.F. Bon, A mechanistic investigation of pickering emulsion
[5] Y. Yang, Z. Fang, X. Chen, W. Zhang, Y. Xie, Y. Chen, Z. Liu, W. Yuan, An overview
polymerization, Polym. Chem. 8 (34) (2017) 5100–5111.
of pickering emulsions: solid-particle materials, classification, morphology,
[34] S.A.F. Bon, P.J. Colver, Pickering miniemulsion polymerization using laponite
and applications, Front. Pharmacol. 8 (2017) 1–20.
clay as a stabilizer, Langmuir 23 (16) (2007) 8316–8322.
[6] P. Wei, Q. Luo, K.J. Edgehouse, C.M. Hemmingsen, B.J. Rodier, E.B. Pentzer, 2D
[35] C. Chiappe, D. Pieraccini, Ionic liquids: solvent properties and organic
Particles at fluid-fluid interfaces: assembly and templating of hybrid structures
reactivity, J. Phys. Org. Chem. 18 (2005) 275–297.
for advanced applications, ACS Appl. Mater. Interf. 10 (26) (2018) 21765–
[36] S.J. Choi, J.W. Won, K.M. Park, P.S. Chang, A new method for determining the
21781.
emulsion stability index by backscattering light detection, J. Food Process Eng.
[7] Q. Luo, P. Wei, Q. Huang, B. Gurkan, E. Pentzer, Carbon capsules of ionic liquid
37 (3) (2014) 229–236.
for enhanced performance of electrochemical double layer capacitors, ACS
[37] Y. Fadil, F. Jasinski, T.W. Guok, S.C. Thickett, H. Minami, P.B. Zetterlund,
Appl. Mater. Interf. 10 (2018) 16707–16714.
Pickering miniemulsion polymerization using graphene oxide: effect of
[8] Q. Luo, P. Wei, E. Pentzer, Hollow microcapsules by stitching together of
addition of a conventional surfactant, Polym. Chem. 9 (2018) 3368–3378.
graphene oxide nanosheets with a di-functional small molecule, Carbon N. Y.
[38] A. Werner, G. Seebe, V. Héroguez, A new strategy to elaborate polymer
106 (2016) 125–131.
composites via pickering emulsion polymerization of a wide range of
[9] P.A. Advincula, A.C. de Leon, B.J. Rodier, J. Kwon, R.C. Advincula, E.B. Pentzer,
monomers, Polym. Chem. 9 (2018) 5043–5050.
Accommodating volume change and imparting thermal conductivity by
278 K. Edgehouse et al. / Journal of Colloid and Interface Science 541 (2019) 269–278

[39] S.D. Kim, W.L. Zhang, H.J. Choi, Pickering emulsion-fabricated polystyrene– [44] Qi Zhang, Chengjian Wang, Fu. Milin, S.Z. Jianli Wang, Pickering High, Internal
graphene oxide microspheres and their electrorheology, J. Mater. Chem. C 2 phase emulsions stabilized by worm-like polymeric nanoaggregates, Polym.
(36) (2014) 7541. Chem. 8 (2017) 5474–5480.
[40] S.H. Che Man, D. Ly, M.R. Whittaker, S.C. Thickett, P.B. Zetterlund, Nano-sized [45] B. Zhao, P. Liu, Y. Jiang, D. Pan, H. Tao, J. Song, T. Fang, W. Xu, Supercapacitor
graphene oxide as sole surfactant in miniemulsion polymerization for performances of thermally reduced graphene oxide, J. Power Sour. 198 (2012)
nanocomposite synthesis: effect of PH and ionic strength, Polym. (United 423–427.
Kingdom) 55 (16) (2014) 3490–3497. [46] X.Y. Peng, X.X. Liu, D. Diamond, K.T. Lau, Synthesis of electrochemically-
[41] R.F.A. Teixeira, H.S. McKenzie, A.A. Boyd, S.A.F. Bon, Pickering emulsion reduced graphene oxide film with controllable size and thickness and its use in
polymerization using laponite clay as stabilizer to prepare armored ‘‘Soft” supercapacitor, Carbon N. Y. 49 (11) (2011) 3488–3496.
polymer latexes, Macromolecules 44 (18) (2011) 7415–7422. [47] T.C. Liu, W.G. Pell, B.E. Conway, Stages in the development of thick cobalt oxide
[42] L. Delafresnaye, P.-Y. Dugas, P.-E. Dufils, I. Chaduc, J. Vinas, M. Lansalot, E. films exhibiting reversible redox behavior and pseudocapacitance,
Bourgeat-Lami, Synthesis of clay-armored poly(vinylidene chloride-co-methyl Electrochim. Acta 44 (17) (1999) 2829–2842.
acrylate) latexes by pickering emulsion polymerization and their film-forming [48] S.C. Petitto, E.M. Marsh, G.A. Carson, M.A. Langell, Cobalt oxide surface
properties, Polym. Chem. 8 (2017) 6217–6232. chemistry: the interaction of CoO(1 0 0), Co3O4(1 1 0) and Co3O4(1 1 1) with
[43] N.R. Cameron, High internal phase emulsion templating as a route to well- oxygen and water, J. Mol. Catal. A Chem. 281 (1–2) (2008) 49–58.
defined porous polymers, Polymer (Guildf). 46 (5) (2005) 1439–1449.

You might also like