You are on page 1of 9

Colloids and Surfaces A: Physicochem. Eng.

Aspects 457 (2014) 49–57

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


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

Influence of a mixed particle/surfactant emulsifier system on


water-in-oil emulsion stability
Alla Nesterenko, Audrey Drelich ∗ , Huiling Lu, Danièle Clausse, Isabelle Pezron
EA 4297 Transformations Intégrées de la Matière Renouvelable UTC/ESCOM, Université de Technologie de Compiègne, rue Personne de Roberval, 60200
Compiègne Cedex, France

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

• Synergic stabilization of W/O


emulsion was allowed using
particles/surfactant system.
• Emulsion stability was reduced with
the increase of surfactant concentra-
tion.
• Maximal emulsion stability was
obtained with a small amount of
surfactant.
• Emulsion was stabilized by particu-
lar 3D network with fine drops and
particles.

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

Article history: Water-in-oil (W/O) emulsions consisting of water and paraffin oil were prepared and stabilized by using
Received 31 March 2014 a dual emulsifier system including hydrophobic silica particles and non-ionic surfactant. The surfac-
Received in revised form 16 May 2014 tant/particles synergistic interaction and more specifically the effect of surfactant concentration (from
Accepted 19 May 2014
0 to 1.8%, w/w) at fixed amount of particles (1.8%, w/w) on the emulsions properties were investigated.
Available online 27 May 2014
Interfacial tension, emulsion morphology, water droplet size distribution, stability with time, rheologi-
cal stress–strain and oscillatory properties were determined to better understand the interaction and
Keywords:
mechanism of emulsion stabilization by solid particles in presence of surfactant. The results showed the
Water-in-oil emulsion
Pickering emulsion
existence of strong interaction between surfactant molecules and silica particles. The critical micellar
Mixed particle/surfactant emulsifier concentration for the mixed-emulsifier system was significantly higher than that of surfactant alone. In
Silica particles presence of a low amount of surfactant (0.1%, w/w), the emulsion was characterized by maximal stability
Rheology (up to 21 days without any phase separation). This dependence was confirmed by rheological measure-
Emulsion stability ments which revealed a highest storage modulus (G ) value for this emulsion. Optical microscopy revealed
the formation of closely packed aggregates of small droplets forming bridges between neighboring large
drops, at low surfactant concentration. The possible explanation for this phenomenon could be the floc-
culation of small emulsion droplets in the presence of low amount of surfactant. The formation of flocs
caused the thickening of the continuous oil phase, the increase in emulsion rigidity and consequently
enhanced emulsion stability.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction

Water-in-oil (W/O) emulsions have different applications in the


food, pharmaceutical, petrochemical and cosmetic industries [1,2].
∗ Corresponding author. Tel.: +33 03 44 23 44 23. In practice, emulsion is a thermodynamically metastable system,
E-mail address: audrey.drelich@utc.fr (A. Drelich). and therefore the main challenge for W/O production is to control

http://dx.doi.org/10.1016/j.colsurfa.2014.05.044
0927-7757/© 2014 Elsevier B.V. All rights reserved.
50 A. Nesterenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 457 (2014) 49–57

the system stability as well as provide its protection against desta- distribution played a key role for the synergic stabilization of W/O
bilization. The separation of the oil and water phases of emulsion emulsion with mixed emulsifier system.
can be observed during storage and/or under effect of temperature Our previous work [5] dealt with W/O emulsion stabilized
and addition of electrolyte compound [3,4]. The common desta- by non-ionic emulsifier or with hydrophobic silica particles. The
bilization mechanisms in emulsion are flocculation, coalescence, results obtained evidenced the difference in the stabilization
sedimentation and Ostwald ripening [5–7]. Fluid W/O emulsions mechanisms between the two systems. Solid particles efficiently
generally present lower stability, compared to oil-in-water (O/W) stabilized the emulsion droplets by forming a network in the
emulsions and can easily coalesce, flocculate or sediment [2]. More- oil phase acting as a steric barrier to coalescence in emulsion.
over, contrary to O/W emulsions stabilized by both steric and After addition of surfactant in particle stabilized W/O emulsion,
electrostatic repulsion, only steric forces are expected to stabilize a progressive destabilization was observed and explained by the
W/O due to the low conductivity of the continuous phase. For the destruction of the particle network by surfactant molecules.
last decade, the better understanding of emulsion stabilization and The challenge of present study is to better understand the
destabilization mechanism has become an increasingly interesting specific interactions between particles and surfactant and their
area for research. effect on emulsion stabilization or destabilization processes. To this
Surfactant molecules with low hydrophile–lipophile balance purpose, W/O emulsions were prepared and stabilized by silica
(HLB), such as polyglycerol polyricinoleate, sorbitan ester surfac- particles solely or mixed with surfactant. The effect of surfactant
tants (Tween and Span series) and soy lecithin, are well known concentration at a fixed amount of particles on emulsion proper-
as emulsifiers used to prepare W/O emulsions [8–11]. More ties was investigated. The evolution of emulsion morphology and
recently, considerable progress has been made in the area of droplet size distribution as function of added emulsifier was fol-
solid-stabilized emulsions (called Pickering emulsions) using sil- lowed by optical microscopy and laser diffraction. The physical
ica, oxides, clay, alumina, carbon, fat crystals and polymer lattices stability of emulsion with time was assessed by visual separation
[12–15]. Furthermore, the use of mixture of surfactant and particles phases observations. Moreover, the rheological measurements pro-
for emulsion stabilization has attracted research and commercial vided the information about viscoelastic and structural properties
attention for the reason of interactions occurring in such systems. of emulsions. The results reported herein contributed to better
However, fewer studies are devoted to the influence of mixed comprehension of W/O stabilization mechanism using dual par-
particle/surfactant emulsifier system in W/O emulsion, and the ticle/surfactant emulsifier system.
mechanism of emulsion stabilization remains poorly understood.
Addition of surfactant molecules to particle system or vice versa
influence their efficiency as emulsifiers [12]. Consequently, the 2. Materials and methods
properties of emulsions obtained are dependent on the syner-
gic interactions between particles and surfactant at the emulsion 2.1. Materials
droplet surface and in the bulk [16,17]. The surfactant has multiple
functions in the stabilization of O/W emulsion in the presence of The paraffin oil characterized by a low viscosity (32 mPa s,
particles: to decrease the interfacial tension favoring droplet for- density of 0.85 g/cm3 ) was purchased from Carl Roth GmbH
mation, to adsorb on the particles surface affecting their wettability (Karlsruhe, Germany). The aqueous phase was high-purity water
and, in some cases, to allow the flocculation of solid particles [18]. with a resistivity of 18.2 M cm produced by a Lab purification
Binks et al. [12] showed that in the presence of constant amount chain (Aquadem/Veolia, Wissous, France). Amorphous fumed sil-
of silica particles, the stability of O/W emulsion was improved ica particles (Aerosil® R711) provided by Evonik Industries AG
at low surfactant (alkylpoly(oxyethylene)) concentrations. In such (Rheinfelden, Germany) were used in this study. Fusion of pri-
system the synergy between particles and surfactant resulted in mary spheres (diameter of 12 nm) into structured aggregates
the emulsion viscosity increase and significant enhancement in the (100–200 nm) during combustion process, led to the formation of
emulsion quality. The fine emulsions of excellent long-term stabil- branched silica network representing the morphology of particles
ity was obtained in the case of surfactant adsorption on particles [5,23]. The partial substitution of silanol groups by methacrylsilane
which was followed by particle agglomeration and adsorption of moieties resulted in the hydrophobic character of Aerosil® R711
such agglomerates at the oil/water interface. With the increase of particles. The non-ionic lipophilic emulsifier (HLB of 4.3), sorbi-
surfactant concentration the displacement of particles by surfac- tan monooelate or Span® 80 was purchased from Sigma–Aldrich
tant from the oil/water interface can occur [19] which resulted in (Buchs, Switzerland).
the particle network breakdown and emulsion destabilization. The
findings from other studies [20,21] demonstrates that the increase
in surfactant concentration at a fixed amount of solid particles led 2.2. Emulsion preparation
to higher cream layer volume which was attributed to the reduction
in O/W emulsion droplet size. This discrepancy between published The external oil phase was prepared by dispersing different
results was probably due to the experimental conditions differ- amounts of Span® 80 (0.1–1.8%, w/w) in paraffin oil and the disper-
ences, but essentially to different variety of surfactant molecules sion was stirred for 15 min at room temperature (20 ± 3 ◦ C) under
and not the same mechanism of stabilization. magnetic stirring at 800 rpm. The fixed amount of hydrophobic sil-
The synergic stabilization mechanism of high internal phase ica particles (1.8%, w/w) was added to paraffin oil (with or without
W/O emulsion was proposed by Zou et al. [22]. Here the authors Span® 80). The obtained mixture of emulsifiers in paraffin oil was
demonstrated that neither surfactant nor particles was suitable homogenized at 24,000 rpm for 3 min using Ultra-Turrax T25 high
to emulsion stabilization. However, the dual emulsifier sys- speed homogenizer equipped with dispersing element S25N-10G
tem allowed satisfactory result for emulsion preparation. In the (Jankel & Kuntel, Staufen, Germany) in order to allow maximum
obtained W/O emulsion, big emulsion droplets were preferentially particle dispersion. The W/O emulsion of 10 g was prepared by
stabilized by particles, while small ones were preferentially sta- adding drop by drop a known volume of water to the obtained
bilized by surfactant molecules. Thus, the increase in particles continuous oil phase. The ingredients were then mixed with the
concentration resulted in higher amount of big emulsion droplets, same homogenizer operating at 24,000 rpm for 5 min. The volume
and the increase in surfactant concentration favored the forma- fraction of water in all emulsions obtained was 30%. Experiments
tion of fine emulsion droplets. Therefore, the wide droplet size were carried out in triplicate.
A. Nesterenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 457 (2014) 49–57 51

2.3. Interfacial tension measurements concentrations of Span® 80 for shear stress values varying from
0.001 to 1000 Pa. The flow curves containing shear rates, shear
Water/oil interfacial tension measurements were carried out by strains and shear viscosities were established. The yield stress val-
use of a K100 Krüss tensiometer (Hamburg, Germany) equipped ues were determined from a plot of shear stress as a function
with a Du Noüy ring geometry. Water/oil interfacial tensions were of shear strain by the tangent crossover method (using two tan-
determined at different concentrations of Span® 80 from 0.001 to gents). Dynamic oscillatory shear measurements were used for the
2% (w/w). To evaluate the silica particles effect on surfactant inter- frequency sweep tests. The viscoelastic linear region was deter-
facial properties, dispersions of surfactant with a fixed amount of mined by preliminary amplitude sweeps (0.01–100% strain, 1 Hz).
1.8% (w/w) particles were prepared following the same procedure Experiment was followed by frequency sweeps ranging from 10
as detailed in the previous section. In this case, dispersions were to 0.1 Hz at a strain of 0.05%. From these measurements, emulsion
first centrifuged (Jouan Centrifuge MR1812, Kirchzarten, Germany) viscoelastic parameters such as storage (G ) and loss (G ) moduli
at 10 000 rpm during 10 min for particles separation. Particles were were determined. All measurements were repeated three times
removed in order to avoid measurement disturbances as particles with each sample.
attachment on the ring and to determine the surface tension of the
surfactant dispersion remaining in the bulk. Only the resulting sur-
3. Results
factant solutions were analyzed. All measurements were repeated
at least three times.
3.1. Emulsion stability with time
2.4. Emulsion characterization
The stability of W/O emulsions as a function of the Span® 80
concentration at constant amount of silica particles (1.8%, w/w)
2.4.1. Stability measurements (bottle test method)
was investigated using bottle test method. As shown in Fig. 1,
The stability of emulsions was studied using the bottle test
the evolution of systems monitored at different time intervals. For
method by observation the sample phase separation with time.
the studied emulsions, the oil phase separation without any water
Freshly prepared emulsions were transferred into 10 mL graduated
phase separation was observed up to 21 days of storage. Experi-
glass bottles sealed with a plastic cap and stored for 21 days at room
mental results clearly show that the phase separation was more
temperature (20 ± 3 ◦ C). The oil phase separation from the emulsion
pronounced at high surfactant concentration. After 21 days, around
was visually monitored at regular time intervals. No aqueous phase
40% of oil was separated in the case of emulsion containing 1.8%
separation was observed for all emulsions after 21 days. The per-
(w/w) of silica particles and 1.8% of surfactant. In contrast to this
centage of each phase volume in relation to the total volume was
result, the improvement in the stability of emulsion in the presence
calculated. Analyses were performed in triplicate.
of small amount of surfactant was noticed. No phase separation was
observed in the case of W/O emulsion stabilized by 1.8% (w/w) of
2.4.2. Morphology and droplet size distribution measurements
silica particles and 0.1% (w/w) of surfactant. Comparing the emul-
The measurements of water droplet diameters and size dis-
sions stabilized by solely silica particles with those stabilized by
tributions in emulsions were performed by laser light diffraction
the mixture of particles and surfactant, it was revealed that in the
instrument (Mastersizer X Malvern Laser Diffraction, Worces-
former case the emulsions had intermediary stability (about 15% of
tershire, UK) equipped with a lens of 100 nm permitting the
oil separated after 21 days) between samples containing low con-
determination of sizes from 0.1 to 200 ␮m. Analyses were realized
centration of Span® 80 and those containing high concentration of
at room temperature (20 ± 3 ◦ C) right after emulsions preparation.
Span® 80. The same tendency was observed for the W/O emulsion
A small quantity (0.05–0.1 mL) of emulsion was introduced in a
stabilized with 1.8% (w/w) of Aerosil® R711 solely in previous study
sample dispersion unit (100 mL) containing the continuous phase
[5]. A slight difference between obtained results could be attributed
(pure paraffin). Diluted emulsions were put into circulation at a
to the difference in the diameter of the dispersing element used for
constant and low flow rate of about 1 mL/min. The typical droplet
emulsion preparation.
size distributions were reported and the volume particle diame-
The stability tests revealed the existence of optimal surfac-
ter (D43 ) was calculated as the mean of three readings per sample.
tant concentration for which a maximal emulsion stability was
For the case where the dispersion of water droplets was incomplete
observed. According to the bottle test results, the more stable emul-
after emulsion dilution, the size of droplet aggregates was detected
sion was obtained using oil phase containing 1.8% (w/w) Aerosil®
by laser diffraction.
R711 and 0.1% (w/w) Span® 80. Moreover, this sample showed a
To complete the information obtained by laser granulometry,
good stability and no phase separation up to 21 days even at 45 ◦ C
the microscopic observation of emulsion morphology was done
(data not shown). To explain this behavior the interfacial proper-
using an optical microscope. A drop of freshly prepared emulsion
ties, emulsion morphology, droplet size and rheological properties
was placed on the glass slide and gently covered with a cover slip.
of mixed emulsifier dispersions as well as W/O emulsions were
The images were processed using a Labophot-2 microscope (Nikon,
analyzed.
Sendai, Japan), linked to a digital video camera (ExwaveHAD, Sony,
Tokyo, Japan) at a magnification of 400×.
3.2. Interfacial properties
2.4.3. Rheological measurements
The rheological flow behavior and viscoelastic properties of In order to evaluate the interaction between silica particles and
samples were measured using a controlled rate rheometer (Phys- surfactant, the interfacial properties of water/paraffin oil system
ica MCR 301, Anton Paar, Graz, Austria) fitted with a cone and were examined. Firstly, to ensure the absence of impurities in the
plate geometry with the diameter of a cone of 50 mm, the cone both oil and water phases, the water/paraffin oil interfacial tension
angle of 1◦ . The temperature was kept at 20 ◦ C during the measure- value of 51.6 ± 0.6 mN/m was determined, as described in previous
ments by a Paar Physica circulating bath and a controlled Peltier study [5]. Then the interfacial tension of water/oil system was mea-
system with accuracy of ±0.1 ◦ C. Samples were carefully placed on sured at different concentrations of Span® 80 in paraffin oil without
the measuring plate and left to rest for 2 min for structure recov- particles or mixed with 1.8% (w/w) of particles. The shapes of the
ery and temperature equilibration. Flow experiments were realized plots of interfacial tension against surfactant concentration are pre-
on the oil phase or on freshly prepared emulsions with different sented in Fig. 2. The results obtained demonstrate that for the
52 A. Nesterenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 457 (2014) 49–57

Fig. 1. Bottle test results demonstrating phase separation as a function of time for W/O emulsions (30/70, v/v) with oil phase containing 1.8% (w/w) of Aerosil® R711 and
different concentrations of Span® 80 surfactant.

dispersions containing only Span® 80, the interfacial tension pro- particles. However, the addition of particles results in an obvi-
gressively decreases with increasing surfactant concentration. The ous shift of the plotted curves into the range of higher surfactant
lowest value of 3.5 mN/m is observed for 0.025% (w/w). As above concentrations. This observation confirms the existence of strong
this concentration the interfacial tension remained constant, it may interactions between surfactant molecules and particles. The sur-
be concluded that the critical micellar concentration (CMC) for the factant molecules would adsorb onto particles having a very large
formation of inverse micelles for Span® 80 is situated between 0.02 specific area [23]. This adsorption results in the decrease of Span®
and 0.03% (w/w). This result is in accordance with previous studies 80 concentration in the bulk and at the water/oil interface. Thus, a
[24–26]. much higher amount of surfactant is required to decrease interfa-
A similar decrease in interfacial tension with increase of Span® cial tension and to come up to the same Span® 80 concentration in
80 concentration is observed for the dispersions containing silica the bulk compared to the system without particles, which explains
the shift phenomenon. As a consequence, the apparent CMC value
in presence of silica particles is significantly higher than that of
sole surfactant system: around 0.5–0.6% (w/w). This adsorption
of surfactant could modify the wettability and the hydrophobic
character of particles therefore altering the properties of formed
emulsion [22,27]. Additionally, the amount of non-adsorbed on
particles surfactant remained in oil dispersion affected the emul-
sification process and emulsion properties, as shown in the next
paragraphs.

3.3. Emulsion morphology and droplet size

To study the emulsion morphology, preparations were ana-


lyzed right after emulsification step by optical microscope. Optical
microscopy images displayed in Fig. 3 show that, at constant par-
ticle amount, the droplet size in W/O emulsions is dependent on
surfactant concentration. An important decrease in water droplet
size is observed with increasing Span® 80 concentrations. This
effect can be associated to the lower water/oil interfacial tension
which reduces the free energy required to create new interface [20].
Fig. 2. Evolution of water/oil interfacial tension with concentration of Span® 80 Thus, during emulsification step the droplet break-up was favored.
surfactant: (䊉) without silica particles and () with 1.8% (w/w) of Aerosil® R711. A similar result was reported earlier [22] in a study dealing with
A. Nesterenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 457 (2014) 49–57 53

Fig. 3. Optical microscopy images of W/O emulsions (30/70, v/v) with oil phase containing 1.8% (w/w) of Aerosil® R711 and different concentrations of Span® 80 surfactant.
Scale bar – 0 ␮m.

W/O emulsion stabilized by silica particles and non-ionic surfac- surfactant remaining in the bulk was insufficient for complete
tant. At the same time, the images demonstrated that emulsion emulsion droplets coverage. Consequently, W/O emulsions stabi-
stabilized by Aerosil® R711 solely has an intermediate droplet size lized with the small amount of surfactant contained both large and
between samples with low and high Span® 80 concentration. fine droplets. Nevertheless, from the concentration of 0.6%, w/w,
Additionally, microscopic observation of the microstructure of the amount of surfactant was enough to cover the totality of water
samples shows that droplets formed have a spherical and well droplets.
defined drop contour, except for samples with low surfactant For the emulsions analyzed, bimodal or trimodal large distribu-
concentration (0.1–0.2%, w/w). For these emulsions, more closely tions were observed indicating the polydisperse character of the
packed aggregates of fine droplets are formed around the large samples, in agreement with the images from optical microscopy.
droplets. This flocculation and aggregation of small droplets allows Moreover, the difference in morphology of emulsions (presence
the creation of protective layer for the big droplets. of flocculated droplets at low surfactant ratio) observed by opti-
Simultaneous measurements of the droplet size distributions cal microscopy affects the mean droplet size. The largest size
in W/O emulsions were realized by laser diffraction. The results of droplets was obtained for the samples with silica particles
obtained confirmed the microscopy observations and showed the and 0.1% (w/w) of surfactant. The mean droplet size varies from
decrease of droplet volume diameter after adding surfactant (Fig. 4). 34.1 ± 1.5 ␮m to 4.4 ± 0.7 ␮m for the emulsions containing 0.1%
This was demonstrated by the displacement of size distributions (w/w) and 1.8% (w/w) of Span® 80, respectively. In the case of W/O
toward lower values as surfactant concentration increased due emulsion stabilized solely by 1.8% (w/w) Aerosil® R711, a mean
to lowering of the interfacial tension. As shown in Fig. 2, at low droplet size of 12.8 ± 0.2 ␮m was obtained. As was shown in our
Span® 80 concentration (0.1–0.2%, w/w), the majority of surfac- previous study [5] the droplet size distribution for the emulsion
tant was adsorbed on silica particles. Considering that the area prepared with 1.8% (w/w) of Span® 80 was centered on around
per molecule of Span® 80 is 32 Å2 /molecule [3], the amount of 1.5 ␮m.

3.4. Rheological characterization

3.4.1. Flow behavior


The rheological behavior of continuous oil phase and W/O emul-
sions with different emulsifier compositions was investigated using
flow experiments to understand the macroscopic mechanism of
stabilization. The amount of Aerosil® R711 silica particles was fixed
at 1.8% (w/w) for all systems studied. The concentration of Span®
80 surfactant was varied from 0 to 1.8% (w/w).
As shown in Fig. 5, the rheological curves obtained for all studied
systems are characterized by a yield stress, contrary to the lin-
ear curve of pure paraffin oil that evidences a Newtonian behavior
[28–30]. Below certain applied stress these samples demonstrated
elastic properties as a solid, but flow like a fluid for threshold above
Fig. 4. Droplet size distributions in volume of W/O emulsions (30/70, v/v) with oil
phase containing 1.8% (w/w) of Aerosil® R711 and different concentrations of Span®
the yield stress value [28,31,32], typical of a gel-like structure. The
80 surfactant. results indicate that the Aerosil® R711 silica particles behave like
54 A. Nesterenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 457 (2014) 49–57

Fig. 5. Stress–strain plots of paraffin oil dispersions (open symbols) and correspond-
ing W/O emulsions (30/70, v/v, filled symbols) with oil phase containing 1.8% (w/w) Fig. 6. Shear rate dependent viscosity profiles of paraffin oil dispersions (open sym-
of Aerosil® R711 (/䊉) without Span® 80, (/) with 0.1% (w/w) of Span® 80 and bols) and corresponding W/O emulsions (30/70, v/v, filled symbols) with oil phase
(/) with 1.8% (w/w) of Span® 80. The solid line corresponds to pure paraffin oil. containing 1.8%, w/w of Aerosil® R711 (/䊉) without Span® 80, (/) with 0.1%
(w/w) of Span® 80 and (/) with 1.8% (w/w) of Span® 80.

a thickening agent of paraffin oil by the formation of a 3D net- whereas loss modulus characterized energy loss due to viscous dis-
work in the bulk phase, as it was suggesting in our previous article sipation [37]. Thus, viscous properties of sample dominated at G
[5]. However, a modification of the rheological profile is observed lower than G and elastic properties dominated at G greater than
depending on the Span® 80 surfactant concentration. G . G (filled symbols) and G (open symbols) of emulsions and
The oil dispersion containing only silica particles is character- paraffin oil dispersion containing 1.8% (w/w) of Aerosil® R711 are
ized by the yield stress of 0.4 ± 0.09 Pa compared to 0.23 ± 0.04 Pa plotted against strain amplitude in Fig. 7.
for the oil dispersion with particles and 0.1% (w/w) of Span® 80 The rheological profiles obtained show that storage and loss
(Fig. 5, open symbols). In agreement with the literature, the addi- moduli are independent of applied strain in low strain range indi-
tion of surfactant can lead to the breakup of silica particle network cating the linear viscoelastic region [29]. G is greater than G for
[19], resulting in a decrease of the yield stress when surfactant all samples studied confirming the viscoelastic solid behavior with
concentration increases, which explains the observed behavior. On gel-like structure. Comparing profiles of studied samples, the low-
the other hand, the rheological profiles obtained for corresponding est moduli values are obtained for oil dispersion containing 1.8%
W/O emulsions are slightly different. The curves obtained for cor- (w/w) Aerosil® R711. Maximal storage modulus values (45–55 Pa)
responding W/O emulsions (Fig. 5, filled symbols) show that the are observed for the emulsion stabilized with particles solely and
dependence of yield stress values is permuted and the emulsion for emulsion stabilized with 1.8% (w/w) of particles and 0.1% (w/w)
stabilized by particles and 0.1% (w/w) of surfactant needs the high- of surfactant. The increase of surfactant amount up to 1.8% (w/w)
est stress (5 ± 0.2 Pa) to flow like a fluid. The emulsion stabilized leads to a low storage modulus value of 20 Pa.
by silica particles solely has an intermediate behavior and a yield Beyond a critical value of strain, a significant decrease of both
stress of 2.3 ± 0.08 Pa. The emulsion containing particles and 1.8% moduli is observed suggesting that the network structure starts
(w/w) of surfactant is characterized by the lowest yield stress of rearranging. At high strain applied, samples showed viscous liquid-
0.37 ± 0.05 Pa. like behavior (G < G ). The transition from solid-like (G > G ) to
The viscosity profiles of oil dispersions (open symbols) and liquid-like (G < G ) behavior indicates the destruction of the net-
corresponding W/O emulsions (filled symbols) with different sur- work [38]. The highest shear strain of 75% needed to break down
factant concentrations at intermediate shear rate domain are given the emulsion structure is observed in the case emulsion containing
in Fig. 6. The results demonstrate that oil dispersions are less silica particles and a low concentration of 0.1% (w/w) of surfactant.
viscous compared to corresponding W/O emulsions. The lowest While for other emulsions, network structure was weaker allow-
viscosity for each system was obtained at 1.8% (w/w) of surfactant. ing an easier rearrangement to accommodate the strain. The lowest
Similarly to Fig. 5, for oil dispersions the sample without surfactant shear strain value of 6% needed for the destruction of emulsion
is the more viscous, whereas for emulsions the sample containing network is obtained at 1.8% (w/w) of particles and high surfactant
0.1% (w/w) of surfactant and 1.8% (w/w) of particles has the high- concentration (1.8%, w/w).
est viscosity. The viscosity of emulsions decreases with increasing In order to investigate the viscoelastic behavior of emulsion,
shear rate, showing their pseudoplastic (or shear-thinning) behav- frequency sweeps were conducted at a fixed strain of 0.05%. The
ior [28,33]. By the application of shear forces, the gel-like structure emulsions studied exhibit the solid-like behavior with G higher
is partially broken and droplets move away from each other. This than G over the entire tested frequency range (Fig. 8). The same
shear-thinning behavior is characteristic for emulsions stabilized observations on the effect of the surfactant concentration are
with solid particles [30,34–36]. noticed.
These rheological results first indicate that the presence of
3.4.2. Viscoelastic properties droplets in emulsion reinforced the network structure of oil dis-
The viscoelastic characteristics of freshly prepared emulsions persions. However, the gel-like structure of emulsions stabilized
were investigated by dynamic oscillatory rheology. From the vari- by silica particles is depending on the Span® 80 concentration. In
ations of storage modulus G and loss modulus G it is possible to agreement with the flow behavior, a high Span® 80 concentration
characterize the strength of network structure of a sample. Stor- weakens the gel-like structure while a low Span® 80 concentra-
age modulus defines the magnitude of energy stored in the sample, tion strengthens it, in comparison to emulsion stabilized solely
A. Nesterenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 457 (2014) 49–57 55

Fig. 7. Strain amplitude dependence of the storage modulus G (filled symbols) and loss modulus G (open symbols) of (♦/) paraffin oil dispersion containing 1.8% (w/w)
of Aerosil® R711 and W/O emulsions (30/70, v/v) with oil phase containing 1.8% (w/w) of Aerosil® R711 (/䊉) without Span® 80, (/) with 0.1% (w/w) of Span® 80 and
(/) with 1.8% (w/w) of Span® 80.

with silica particles. Moreover, the rheological characterizations


are in good concordance with stability tests showing an interme-
diate stability of emulsion without surfactant (stabilized by 1.8%,
w/w, of silica particles) between samples with low and high Span®
80 concentrations.

4. Discussion

Surprisingly, the different results show that a low amount of


Span® 80 surfactant seems to improve the stability of emulsion
stabilized by Aerosil® R711 silica particles. It is assumed that, in
the absence of surfactant, silica particles create a steric protective
shield for emulsion droplets [39]. Moreover, due to strong cohesive
interaction between silica particles, agglomerates of particles lead
to the formation of a 3D network in the continuous phase, which
offer a good stability to emulsion [5].
Once an amount of Span® 80 is added, interfacial mea-
surements show that surfactant molecules are adsorbed on the
particle surface. At high Span® 80 concentration, enough surfactant
molecules may screen the cohesive interactions between silica par-
ticles allowing the particle network redispersion [12]. The results
Fig. 8. Frequency dependence of the storage modulus G (filled symbols) and loss
obtained by rheological measurement confirmed this tendency and
modulus G (open symbols) of W/O emulsions (30/70, v/v) with oil phase containing
1.8% (w/w) of Aerosil® R711 (/䊉) without Span® 80, (/) with 0.1% (w/w) of showed a weak association of emulsion system stabilized by 1.8%
Span® 80 and (/) with 1.8% (w/w) of Span® 80. (w/w) of Aerosil® R711 and 1.8% (w/w) of Span® 80. The change
56 A. Nesterenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 457 (2014) 49–57

in the 3D particles network structure, in presence of Span80, is the the emulsion structure was particularly high, confirming the strong
result of the sedimentation of the emulsion and the separation of association of droplets in emulsion. A possible explanation for this
the oil phase. Pichot et al. [40] note that at high surfactant concen- phenomenon could be the formation of particular stabilizing net-
trations, the structure of mixed-emulsifier emulsions is very similar work in the emulsion using both fine droplets and silica particles.
to that of emulsions stabilized by surfactant alone. Dutschk et al. Such network structure allowed thickening of the oil continuous
[14] affirmed that in presence of solid particles surfactant can act as phase improving the strength of emulsifier tridimensional network
a plasticizing agent increasing the fluidity of emulsion and reducing structure and stability of emulsion. The results of this paper con-
the strength of the 3D network between droplets. tributed to a better understanding of W/O emulsion stabilization
In the case of a small addition of Span® 80 surfactant, a compet- mechanism using mixed particle/surfactant emulsifier system.
itive adsorption of the surfactant molecules exist on the particles
surface and at the W/O droplet interface. The presence of surfactant Acknowledgements
molecules favors the formation of small droplets during emulsifi-
cation process. But at the same time, the interfacial tension was The European Space Agency is gratefully acknowledged for the
remained high, and relatively large droplets are also formed in financial support from the MAP-projects FASES (Fundamental and
emulsion. As shown in optical microscopy images, the fine droplets Applied Studies of Emulsion Stability) and PASTA (Particle Stabi-
can flocculate and form bridges between neighboring large drops. lized Emulsions and Foams).
Consequently, the stabilizing network in the emulsion is formed
by both fine droplets and silica particles. Proposed explanation
References
is in very good agreement with the results reported by Zou et
al. [22]. Authors confirmed that the polydisperse droplet distri- [1] S. Ghosh, D. Rousseau, Fat crystals and water-in-oil emulsion stability, Curr.
bution and the simultaneous presence of big and fine droplets Opin. Colloid Interface Sci. 16 (2011) 421–431.
are particularly important to allow a long-term stability of W/O [2] F.Y. Ushikubo, R.L. Cunha, Stability mechanisms of liquid water-in-oil emul-
sions, Food Hydrocolloids 34 (2014) 145–153.
emulsions stabilized by dual particle/surfactant emulsifier sys- [3] F.O. Opawale, D.J. Burgess, Influence of interfacial properties of lipophilic sur-
tem. The formation of such network structure caused thickening factants on water-in-oil emulsion stability, J. Colloid Interface Sci. 197 (1998)
of the continuous phase, provided enhanced resistance to coales- 142–150.
[4] D. Clausse, I. Pezron, L. Komunjer, Stability of W/O and W/O/W emulsions as
cence of emulsion and resulted in its excellent long-term stability.
a result of partial solidification, Colloid Surf. A: Physicochem. Eng. Aspect 152
Macierzanka and Szelag [41] demonstrated that rheological prop- (1999) 23–29.
erties of W/O depended on the emulsion droplet size. Large droplet [5] A. Drelich, F. Gomez, D. Clausse, I. Pezron, Evolution of water-in-oil emulsions
stabilized with solid particles. Influence of added emulsifier, Colloid Surf. A:
emulsion contained higher amount of non-adsorbed solid particles
Physicochem. Eng. Aspect 365 (2010) 171–177.
acting as shear-thinning agent which resulted in higher emulsion [6] D.J. McClements, Edible nanoemulsions: fabrication, properties, and functional
yield stress. performance, Soft Matter 7 (2011) 2297–2316.
Comparing to our previous results dealing with W/O emulsion [7] T.F. Tadros, Emulsion Science and Technology, WILEY-VCH Verlag GmbH & Co,
Weinheim, 2009.
stabilized by Aerosil® R711 or Span® 80 [5], present study suggests [8] L. Ambrosone, M. Mosca, A. Ceglie, Impact of edible surfactants on the oxidation
that for the maximal emulsion stability the use of combination of of olive oil in water-in-oil emulsions, Food Hydrocolloids 21 (2007) 1163–1171.
particles and small amount of surfactant was the most preferred [9] I. Gülseren, M. Corredig, Interactions between polyglycerol polyricinoleate
(PGPR) and pectins at the oilewater interface and their influence on the stability
compared to each emulsifier alone. The dual emulsifier system of water-in-oil emulsions, Food Hydrocolloids 34 (2014) 154–160.
allowed the synergic stabilization of emulsion. The research of [10] M. Porrasa, C. Solans, C. Gonzalez, G.M. Gutierrez, Properties of water-in-oil
Binks et al. [12] confirmed a crucial role of the synergism between (W/O) nano-emulsions prepared by a low-energy emulsification method, Col-
loid Surf. A: Physicochem. Eng. Aspect 324 (2008) 181–188.
surfactant and particles in the emulsion stabilization. Only in the [11] L.C.B. Züge, C.W.I. Haminiuk, G.M. Maciel, J.L.M. Silveira, A.P. Scheer, Cata-
case of mixed emulsifier system with particles and surfactant a strophic inversion and rheological behavior in soy lecithin and Tween 80 based
suitable quality of emulsions in terms of stability was obtained. food emulsions, J. Food Eng. 116 (2013) 72–77.
[12] B.P. Binks, A. Desforges, D.G. Duff, Synergistic stabilization of emulsions by a
mixture of surface-active nanoparticles and surfactant, Langmuir 23 (2007)
1098–1106.
5. Conclusion [13] S. Haj-shafiei, S. Ghosh, D. Rousseau, Kinetic stability and rheology of wax-
stabilized water-in-oil emulsions at different water cuts, J. Colloid Interface
Sci. 410 (2013) 11–20.
This study developed the use of double emulsifier system con- [14] V. Dutschk, J. Chen, G. Petzold, R. Vogel, D. Classe, F. Ravera, L. Liggieri, The role of
taining silica particles Aerosil® R711 and non-ionic surfactant emulsifier in stabilization of emulsions containing colloidal alumina particles,
Span® 80 for water-in-oil (W/O) emulsion stabilization. The influ- Colloid Surf. A: Physicochem. Eng. Aspect 413 (2012) 239–247.
[15] F. Gautier, M. Destribats, R. Perrier-Cornet, J.F. Dechézelles, J. Giermanska, V.
ence of surfactant concentration at constant amount of particles on Héroguez, S. Ravaine, F. Leal-Calderon, V. Schmitt, Pickering emulsions with
emulsions properties was particularly investigated. The existence stimulable particles: from highly- to weakly-covered interfaces, Phys. Chem.
of interactions between particles and emulsifier was evidenced by Chem. Phys. 9 (2007) 6455–6462.
[16] C.P. Whitby, D. Fornasiero, J. Ralston, Effect of oil soluble surfactant in emulsions
interfacial tension measurements. As a result, the CMC value for stabilised by clay particles, J. Colloid Interface Sci. 323 (2008) 410–419.
mixed-emulsifier system was significantly increased compared to [17] H. Nciri, N. Huang, V. Rosilio, M. Trabelsi-Ayadi, M. Benna-Zayani, J.L. Grossiord,
surfactant alone. The raise in surfactant concentration resulted in Rheological studies in the bulk and at the interface of Pickering oil/water emul-
sions, Rheol. Acta 49 (2010) 961–969.
the reduced emulsion droplet size which was revealed by opti- [18] B.R. Midmore, Synergy between silica and polyoxyethylene surfactants in the
cal microscopy and laser diffraction measurements. Moreover, the formation of O/W emulsions, Colloid Surf. A: Physicochem. Eng. Aspect 145
presence of closely packed aggregates around the large droplets (1998) 133–143.
[19] C. Vashisth, C.P. Whitby, D. Fornasiero, J. Ralston, Interfacial displacement of
was observed in the case of emulsion stabilized by small amount
nanoparticles by surfactant molecules in emulsions, J. Colloid Interface Sci. 349
of surfactant (0.1%, w/w) and particles. This emulsion remained (2010) 537–543.
with highest stability up to 21 days without any phase separation. [20] R. Pichot, F. Spyropoulos, I.T. Norton, Mixed-emulsifier stabilised emulsions:
Investigation of the effect of monoolein and hydrophilic silica particle mixtures
While the emulsions with 0 or 1.8%, w/w of surfactant and par-
on the stability against coalescence, J. Colloid Interface Sci. 329 (2009) 284–291.
ticles showed significantly lower stability. Thus, an optimal W/O [21] J. Wang, F. Yang, J. Tan, G. Liu, J. Xu, D. Sun, Pickering emulsions stabilized by a
emulsion stabilization was achieved at low surfactant concentra- lipophilic surfactant and hydrophilic plate-like particles, Langmuir 26 (2010)
tion. The oscillatory tests showed that W/O emulsions exhibited 5397–5404.
[22] S. Zou, Y. Yang, H. Liu, C. Wang, Synergistic stabilization and tunable structures
solid-like behavior over the entire tested frequency range. At 0.1% of Pickering high internal phase emulsions by nanoparticles and surfactants,
(w/w) of surfactant, the shear strain value needed to break down Colloid Surf. A: Physicochem. Eng. Aspect 436 (2013) 1–9.
A. Nesterenko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 457 (2014) 49–57 57

[23] L. Forny, I. Pezron, K. Saleh, P. Guigon, L. Komunjer, Storing water in powder rupture rheological characterization of fluid gels), Les Cahiers de Rhéol. 16
form by self-assembling hydrophobic silica nanoparticles, Powder Technol. 171 (1998) 46–53.
(2007) 15–24. [33] S.M. Hanning, T. Yu, D.S. Jones, G.P. Andrews, J.A. Kieser, N.J. Medli-
[24] E. Santini, L. Liggieri, L. Sacca, D. Clausse, F. Ravera, Interfacial rheology of Span cott, Lecithin-based emulsions for potential use as saliva substitutes in
80 adsorbed layers at paraffin oil–water interface and correlation with the cor- patients with xerostomia – viscoelastic properties, Int. J. Pharm. 456 (2013)
responding emulsion properties, Colloid Surf. A: Physicochem. Eng. Aspect 309 560–568.
(2007) 270–279. [34] H. Huang, B. You, S. Zhou, L. Wu, Rheological behavior of aqueous organosilicone
[25] W. Wang, K. Li, P. Wang, S. Hao, J. Gong, Effect of interfacial dilational rheology resin emulsion stabilized by colloidal nanosilica particles, J. Colloid Interface
on the breakage of dispersed droplets in a dilute oil–water emulsion, Colloid Sci. 310 (2007) 121–127.
Surf. A: Physicochem. Eng. Aspect 441 (2014) 43–50. [35] M.V. Tzoumaki, T. Moschakis, V. Kiosseoglou, C.G. Biliaderis, Oil-in-water emul-
[26] L.J. Peltonen, J. Yliruusi, Surface pressure, hysteresis, interfacial tension, and sions stabilized by chitin nanocrystal particles, Food Hydrocolloids 25 (2011)
CMC of four sorbitan monoesters at water–air, water–hexane, and hexane–air 1521–1529.
interfaces, J. Colloid Interface Sci. 227 (2000) 1–6. [36] B. Yaghi, Rheology of oil-in-water emulsions containing fine particles, J. Petrol.
[27] A. Kawazoe, M. Kawaguchi, Characterization of silicone oil emulsions stabi- Sci. Eng. 40 (2003) 103–110.
lized by TiO2 suspensions pre-adsorbed SDS, Colloid Surf. A: Physicochem. Eng. [37] D. Saha, S. Bhattacharya, Hydrocolloids as thickening and gelling agents in food:
Aspect 392 (2011) 273–287. a critical review, J. Food Sci. Technol. 47 (2010) 587–597.
[28] B. Niraula, T.C. King, M. Misran, Rheology properties of dodecyl-␤-d-maltoside [38] R. Shu, W. Sun, T. Wang, C. Wang, X. Liu, Z. Tong, Linear and nonlinear vis-
stabilized mineral oil-in-water emulsions, Colloid Surf. A: Physicochem. Eng. coelasticity of water-in-oil emulsions: effect of droplet elasticity, Colloid Surf.
Aspect 231 (2003) 159–172. A: Physicochem. Eng. Aspect 434 (2013) 220–228.
[29] T. Tadros, Application of rheology for assessment and prediction of the long- [39] Y. Chevalier, M.A. Bolzinger, Emulsions stabilized with solid nanoparticles:
term physical stability of emulsions, Adv. Colloid Interface Sci. 108/109 (2004) Pickering emulsions, Colloid Surf. A: Physicochem. Eng. Aspect 439 (2013)
227–258. 23–34.
[30] L.G. Torres, R. Iturbe, M.J. Snowden, B.Z. Chowdhry, S.A. Leharne, Preparation of [40] R. Pichot, F. Spyropoulos, I.T. Norton, O/W emulsions stabilised by both
o/w emulsions stabilized by solid particles and their characterization by oscil- low molecular weight surfactants and colloidal particles: the effect of
latory rheology, Colloid Surf. A: Physicochem. Eng. Aspect 302 (2007) 439–448. surfactant type and concentration, J. Colloid Interface Sci. 352 (2010)
[31] R.P. Chhabra, Bubbles, Drops and Particles in Non-Newtonian Fluids, CRC Press, 128–135.
Boca Raton, FL, 2006. [41] A. Macierzanka, H. Szelag, Microstructural behavior of water-in-oil emulsions
[32] C. Michon, G. Cuvelier, E. Aubree, B. Launay, Caractérisation rhéologique stabilized by fatty acid esters of propylene glycol and zinc fatty acid salts,
de gels fluides aux petites déformations et à la rupture (Small strains and Colloid Surf. A: Physicochem. Eng. Aspect 281 (2006) 125–137.

You might also like