Você está na página 1de 11

Colloids and Surfaces A: Physicochem. Eng.

Aspects 508 (2016) 316–326

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


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

Bulk and interfacial rheology of emulsions stabilized with clay


particles
Joung Sook Hong a,∗,1 , Peter Fischer b
a
Department of Chemical Engineering, Soongsil University, Seoul 156-743, Republic of Korea
b
Institute of Food, Nutrition and Health, ETH Zurich, 8092 Zurich, Switzerland

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

h i g h l i g h t s

• Interfacial aggregation of natural/modified clays causes high interfacial modulus.


• High interfacial modulus induces significant emulsion stabilization.
• Low interfacial modulus of natural clays does not induce stabilization.
• High modulus of the bulk oil phase with OMt clays causes stabilization.

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

Article history: This contribution aims to investigate the stabilizing effect of clay particles on oil-in-water emulsions
Received 10 April 2016 using bulk and interfacial rheological measurements. Depending on the surface properties of clay parti-
Received in revised form 27 July 2016 cles, the emulsion showed different rheological responses as a result of different stabilization behavior.
Accepted 22 August 2016
Hydrophilic clays (natural montmorillonite (NMt)) stabilized the emulsion via an interfacial layer with
Available online 23 August 2016
distinguished interfacial modulus (G*interface ), while hydrophobic clays (organically modified montmo-
rillonite (OMt)) stabilized the emulsion by increasing the bulk modulus of the oil phase (G*oil ). In
Keywords:
composite systems, when NMt is present together with cationic surfactant or OMt, complex forma-
Interfacial rheology
Clays
tion at the oil-water interface significantly influenced the rheological response and emulsion stability.

∗ Corresponding author.
E-mail addresses: polymer@ssu.ac.kr, jsook.hong@gmail.com (J.S. Hong).
1
Currently address: Department of Chemical Engineering and Materials Science, Michigan State University, East Lansing, Michigan 48824, USA.

http://dx.doi.org/10.1016/j.colsurfa.2016.08.040
0927-7757/© 2016 Elsevier B.V. All rights reserved.
J.S. Hong, P. Fischer / Colloids and Surfaces A: Physicochem. Eng. Aspects 508 (2016) 316–326 317

Aggregation The high interfacial modulus (G*interface ) of the composite interface stabilized the flocculated droplets more
Interfacial location than the repulsive interaction between droplets. As a consequence, the bulk modulus of the emulsion
Emulsion (G*emulsion ) also increases. Finally it could be shown that with increasing interfacial area (oil-to-water
Stabilization ratio), microscopic and macroscopic stabilization of emulsions could be improved. Rheological study
provided structural information of the complex interface and defined the role of clay particles for the
stabilization.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction relate the diverse mechanical response to the formation of an inter-


facial adsorption layer [10,11,24,25,29]. This in particular, because
Emulsions composed of immiscible phases are thermodynami- complex physical and chemical interaction compete during the
cally unstable. They continuously change structure to minimize the build-up of an interfacial layer, rapid diffusion of surfactants, as well
free energy. Against this nature of emulsion, during emulsification as addition of salt to the interface or colloidal particles enhances
both, dispersed and continuous phase develop structural domains precipitation of particles [13] or interfacial accumulation [30].
depending on composition and due to the external energy input. Despite such remarkable efforts to measure interfacial proper-
The competition of breakup and coalescence results in heteroge- ties, we are far from knowing how strong an interfacial modulus
neous size distribution of droplets. After the emulsifying process, (G*interface ) has to be to provide emulsion stability against the coa-
coalescence between droplets continues and the emulsion finally lescence induced by the density difference between the oil and
separates into its individual phases. Coalescence is a thermody- aqueous phases or by flocculation of droplets [19,31–34]. Addi-
namically favorable process driven by net reduction of interfacial tionally, it is difficult to predict how the bulk properties of the
energy and results in film drainage and rupture of the continu- emulsion are influenced by the interfacial response of the emul-
ous phase film between the dispersed droplets [1,2]. Stabilization sion system after processing. Because the stabilization is obtained
against coalescence is one of the most important parameters to as a result of an integrated association of microscopic interfacial
determine product quality of the emulsion. Thus a large number phenomena of particles and build-up of macroscopic inter-droplets
of publications focused on the design of stable macro- or micro- structure, stabilization of the complicated emulsion system cannot
emulsions. For stabilization, small molecular weight surfactants, be obtained if either micro- or macrostructural contributions are
block-copolymers, proteins, colloidal particles, and combinations neglected. Therefore, for successful application of particles for the
thereof are added to the emulsions [3–14]. All mentioned materials stabilization of the emulsions, it is necessary to figure out the micro-
form an interfacial adsorption layer depending on concentration, scopic/macroscopic motion of particles in the emulsion system
hydrodynamic condition, and physicochemical properties such under flow conditions and define their role for emulsion stabiliza-
as hydrophilic-lipophilic balance and contact angle. As a result tion. There are still many challenges remaining to understand the
the dispersed droplets are preserved from coalescence by sev- interfacial phenomena of particles existing in nature and define
eral mechanisms such as electrostatic repulsion, interfacial tension their stabilization ability.
reduction, and Marangoni stress against film drainage [15–19]. The In this study, the emulsion stabilization effect of clay parti-
combination of surfactant and particles expects to give a synergic cles was studied due to their abundance in nature, high potential
stabilization effect to emulsions by a formation of electrostatically in many applications such as industrial materials, water treat-
and mechanically strengthened interface. The head or tail of the ment [35], biomedical usage [36], and industrial needs to improve
ionic surfactants adsorbs on the surface of particles by electrostatic processing efficiency in fluid transport [37,38]. Especially, mont-
attraction and renders the surface more hydrophobic or hydrophilic morillonite, a member of the smectite group of clay minerals
[3]. have been studied as a colloidal particle in diverse applications.
In order to characterize the interfacial adsorption layer, sev- For example, variations of montmorillonite were developed by
eral mechanical methods have been developed in the past [20–22]. exchange of an alkaline metal existing in the interlayer with organic
Since this frontier work in interfacial rheology, the interfacial molecules. For the purpose of understanding their behavior at
layer built-up by particles or surfactants is acknowledged to the interface and their influence on the emulsion stabilization,
have independent mechanical properties than the surrounding two montmorillonite samples of different surface properties were
bulk phases. Research on the formation of an interfacial layer of chosen and added to an emulsion system. Emulsions with clay
surfactants, amphiphilic molecules, and particles by using dilata- particles, particle mixtures and clay-surfactant blends were inves-
tional and shear viscosity measurements contributed to further tigated based on the relationship between rheological response and
understanding of their adsorption kinetics and stabilization effect structure of emulsion systems. Emulsion rheology, interfacial rhe-
[11,19,23–26]. ology as well as morphology were expected to provide important
Interfacial adsorption of surface-active materials increases information to suggest the stabilization mechanism of clays in the
interfacial modulus (G*interface ) depending on concentration as emulsion system.
expected. However, it takes place over a long time to reach a stable
interfacial structure because interfacial structures keep changing 2. Materials and methods
due to the change of several external conditions such as electrolyte,
temperature, and flow. This structural rearrangement is reflected 2.1. Materials
in the interfacial modulus (G*interface ). With particles, the interfa-
cial modulus is expected to be high enough to induce stability no For oil-in water emulsions, canola oil (CJ CheilJedang Co., Korea)
matter how fast they build the interfacial layer [2,13,27,28]. Never- was used for the oil phase, which mostly consists of unsatu-
theless particles showed unpredictable interfacial layer formation, rated fatty acid (91.3%). The oil had a density of 0.85 g/cm3 and
especially with surfactant or salt. For complex systems including a viscosity of 0.06 Pas at 20 ◦ C. For the water phase, purified
particles, surfactants, or salt, interfacial rheology combined with (Barnstead RO pure LP system, Thermo Scientific, USA) and deion-
diverse rheological measurement techniques was suggested to ized (Nanopure II systems, Thermo Scientific, USA) water was
318 J.S. Hong, P. Fischer / Colloids and Surfaces A: Physicochem. Eng. Aspects 508 (2016) 316–326

used. Clay samples were purchased from Southern Clay Prod- MCR 501 rheometer equipped with biconical geometry [25]. Oscil-
ucts Inc. (USA). Two different clay samples were used: natural latory amplitude sweep experiments (␻ = 1 s−1 ) were performed
Na-montmorillonite and organically modified montmorillonite. as a function of strain (␥) from 0.01 to 10% and dynamic frequency
Natural Na-montmorillonite (Cloisite® Na+) (NMt) had a density sweep experiments (␥ = 0.5%) were performed over the angular fre-
of 2.6 g/cm3 . When NMt is organically modified with a surfactant quency (␻) from 0.05 to 10 rad/s. The transient evolution of the
with a long hydrocarbon tail (C16–18 ), the clay surface had more interfacial adsorption layer was investigated by an oscillatory time
hydrophobic characteristics. The organically modified montmo- sweep for 24 h (␥ = 0.1%, ␻ = 1 s−1 ). Temperature was maintained
rillonite (OMt) was a dimethyl hydrogenated-tallow ammonium for all experiments at 20 ◦ C.
modified montmorillonite (Cloisite® 20A), which was prepared by
exchanging ions in NMt with alkyl ammonium cations. Both clay
samples were added to the emulsion as received. In the emulsion 3. Results
systems, the size of clays was distributed from several hundreds
of nanometers to several micrometers depending on the degree of 3.1. Pure clay adsorption layers
exfoliation and aggregation (supplementary Fig. 1).
The lack of thermodynamic miscibility between oil and water
2.2. Preparation of emulsions continuously drives an oil-in-water emulsion into phase separation
if there is no external mixing energy. Such a continuous struc-
To prepare clay stabilized emulsion, water, clay particles, and oil tural change makes it difficult to measure precise and reproducible
were sequentially added to vessel. The amount of oil and water was rheological properties of the emulsions to provide structural infor-
fixed to 5 g and the mixing ratio of oil and water was varied over a mation even though measurements are carried out consistently
wide range from 30/70 wt/wt to 70/30 wt/wt oil-in-water (O/W). and repetitively. The oil-in-water emulsion system tested in this
Plus, an additional mass of clay samples (wtclays /wtoil+water ) were study also never showed stability against phase separation over
added to oil and water (5 g). For example, to prepare 40/60 wt/wt the whole composition region. For this oil-in-water emulsion sys-
O/W emulsions with 3000 ppm clay, oil of 2 g, clay of 15 mg, and tem, two clay samples with different surface property were added
water of 3 g were added to the vessel and emulsified all together. to investigate the effect of clays for the emulsion stabilization.
The emulsification was performed using a vortex mixer (MX-S, According to our previous study, clays showed different interfa-
PRONEER Co., Korea) under fixed stirring conditions of 1600 rpm cial localization depending on their wetting behavior between oil
for 10 min at room temperature. Right after emulsification, the and water [39] and also they reduced interfacial tension in dif-
emulsion was left under vacuum to remove bubbles (700 mmHg, ferent way [40]. As shown in Fig. 1, with the addition of clays
5 min) and the separated aqueous or oil phase was not removed. (3000 ppm), the emulsions showed different stabilizing behavior
Pure emulsions without clays were prepared in the same way. depending on emulsion composition and clay type (more details
In this study, the emulsion stability was mentioned based on in supplementary Fig. 2–6). Clay particles improved the emulsion
macroscopic/microscopic observations for 30 min after emulsify- stability against phase separation. Especially, as the amount of the
ing (supplementary Fig. 2–6) because rheological measurement oil phase increases (increased O/W ratio), the emulsions with OMt
was finished in 30 min after emulsifying. From the preliminary were more stable than those stabilized with NMt.
stabilization test over a wide range of emulsion compositions (sup- In order to investigate the stabilization effect of clay particles,
plementary Fig. 2–6), the 40/60 wt/wt O/W emulsion was used as the composition of the emulsion was fixed at 40/60 wt/wt O/W.
the model emulsion because it was the most stable sample. The concentration of clay varied from 3000 to 20,000 ppm for both
clay samples. The 40/60 wt/wt O/W emulsion showed average drop
2.3. Optical observations size of 80 ␮m with NMt (3000 ppm), while it was 160 ␮m with OMt
(3000 ppm) (supplementary Fig. 6). As time goes by, the emulsion
After the preparation of the emulsions, they were studied for stabilized with NMt showed separation of the water phase at the
macroscopic and microscopic changes. For macroscopic observa- bottom, while the emulsion stabilized with OMt showed an oil layer
tions, images (2560 × 1712 pixels) of the vessel containing the at the top of the emulsion. Even with higher concentration of clay,
emulsion were taken every 5 min to monitor possible phase sep- the emulsions have the same tendency toward phase separation
aration. To address microscopic changes, small amounts of the (supplementary Fig. 3). Prior to analyzing the rheological data, the
emulsions were transferred onto a glass slide. The space between stability of the emulsion samples were macroscopically checked
the glass slide and cover glass was consistently set at 180 ␮m. Pic- immediately after the rheological test (∼10 min from loading to
tures were taken (1280 × 1024 pixels, Olympus BX43, 4×, 10×, and measuring) to ensure the reliability of the test (Fig. 2). In the case of
20 x lens, Japan) to observe the structure of the emulsion. the 40/60 wt/wt O/W emulsion without clays, the emulsion quickly
started separation after emulsifying. During the measurement, it
2.4. Rheological measurements was completely separated into two phases. As shown in Fig. 2(a),
the water phase stayed in the center and the oil phase spread out-
For rheological measurements of the emulsions, a sample of side. When the emulsion was mixed with clays, the stability of
constant volume was used following a predefined procedure to the emulsions was significantly improved even though the emul-
minimize the effect of emulsion aging. After emulsifying, the emul- sions still showed partial phase separation. With NMt (6000 ppm)
sion had a bubble layer on the top. To exclude bubbles, the emulsion (Fig. 2(b)), the emulsion separated only mildly during the measur-
sample was taken from the middle part of the sample. A dynamic ing period. The emulsion with OMt (6000 ppm) remained stable
oscillatory sweep tests were performed using a rheometer (AR-G2, during the rheological measurement (Fig. 2(c)). These macroscopic
TA instruments, USA) with a parallel plate fixture (60 mm in diam- observations indicated that clay as a colloidal particle stabilized the
eter) to measure the complex viscosity (␩*(Pas)), storage modulus emulsion in different ways depending on the surface properties of
(G’(Pa)), and loss modulus (G”(Pa)) as a function of strain and fre- the clay.
quency. Prior to frequency sweep test, oscillatory strain sweep test The different stabilizing behavior of both clay samples was
was performed to confirm the linear viscoelastic condition of strain. reflected in the rheological response of the emulsions. The linear
The strain was fixed to 5% for frequency sweep test. The interfa- viscoelasticity response of 40/60 wt/wt O/W emulsions was com-
cial shear rheology was performed using an Anton Paar Physica pared depending on the concentration of two clay samples (Fig. 3).
J.S. Hong, P. Fischer / Colloids and Surfaces A: Physicochem. Eng. Aspects 508 (2016) 316–326 319

Fig. 1. Comparison of the emulsion stability depending on emulsion composition and clay type. The images were taken at 30 min after emulsifying. The outer diameter of
vial is 28 mm.

Fig. 2. Optical observation of the emulsion samples after rheological measurement: (a) the 40/60 wt/wt O/W emulsion, (b) the 40/60 wt/wt O/W emulsion with NMt 6000 ppm,
and (c) the 40/60 wt/wt O/W emulsion with OMt 6000 ppm. When the rheological measurement test completed, the upper fixture of measuring geometry was lifted to inspect
the emulsion samples left on the bottom fixture. Also the emulsions remained in the vessel were cross-checked.

Fig. 3. Oscillatory frequency sweep test of the 40/60 wt/wt O/W emulsions depending on the concentration of both NMt and OMt: (a) emulsion storage modulus (G’emulsion )
and (b) emulsion loss modulus (G”emulsion ).
320 J.S. Hong, P. Fischer / Colloids and Surfaces A: Physicochem. Eng. Aspects 508 (2016) 316–326

Fig. 5. Oscillatory time sweep test of interfacial storage and loss modulus (G’interface
and G”interface ) of the oil-water interface with adsorbed NMt and OMt clay particles.
Fig. 4. Oscillatory frequency sweep test showing storage and loss moduli (G’oil and
G”oil ) of the oil phase depending on the concentration of OMt.
oil phase dispersed with OMt increased its modulus (G*oil ) signif-
icantly because OMt dispersed well due to the chemical affinity
With NMt, the emulsions were not stable regardless of the concen- between OMt and the oil. The oscillatory frequency sweep test of
tration of NMt as shown in Fig. 1 and Fig. 2 (b) (supplementary Fig. the oil phase shows the development of a plateau in the storage
2). Storage modulus of the emulsions (G’emulsion ) with NMt showed modulus (G*oil ) for low frequencies with increasing amounts of OMt
a plateau behavior in the lower frequency region. When the con- as shown for 3000 and 6000 ppm in Fig. 4.
centration of NMt was increased from 3000 ppm to 6000 ppm, the A significant increase in the modulus of the oil phase (G*oil )
emulsion modulus (G*emulsion ) significantly increased. When more also increases in the modulus of the emulsion (G*emulsion ), might
NMt was added, the emulsion modulus (G*emulsion ) decreased. The delay coalescence and, thus, results in a stable emulsions. This sta-
rheological behavior was changed from a gel-like behavior (with bilization behavior could be observed only when the composition of
6000 ppm) to a pregel-like behavior (with 20,000 ppm). It means the oil phase was higher than 40%. Bulk rheology of the emulsions
that the emulsion changes its behavior depending on the con- showed that both clay samples with different surface properties
centration of NMt. Initially, NMt was expected to localize at the have different stabilizing potential depending on the concentra-
interface by its balanced wetting behavior [39]. If clays kept accu- tion of the clay. To specify their interfacial behavior, interfacial
mulating in the interface, the modulus of the emulsion (G*emulsion ) shear rheology of the oil-water interface with both clay samples
had to be increased or stayed constant with an increasing con- was performed. The oscillatory time sweep test of the interface
centration of clays because the addition of NMt slightly changed was carried out to observe long-term interfacial behavior (Fig. 5).
interfacial tension of the O/W emulsion with the increasing con- The interface with NMt showed an increasing storage interfacial
centration of clays [40], which resulted in slight increase in the storage modulus (G’interface ) due to the continuous sedimentation
interfacial area of the emulsion. However, the emulsion modu- of dispersed clay particles to the interface. However, the interfa-
lus (G*emulsion ) instead decreased with increasing concentration of cial moduli (G’interface, and G”interface ) were very low ( « 0.01 Pa m).
NMt. It means that concentrated NMt tends to desorb to the water Adsorption and interfacial structure of the NMt layer was easily dis-
phase rather than accumulate in the interface. Therefore, the emul- rupted by shear force and gravity. The stabilization effect of NMt
sion modulus (G*emulsion ) did not increase with the concentration also could not be improved by increased concentration of NMt. The
of NMt and the emulsion could not remain stable even though the oil-water interface with OMt did not respond to the interfacial rhe-
concentration of NMt was increased. ology tests except displaying a minor loss moduli (G”interface ) [39].
On the other hand, the emulsion modulus (G*emulsion ) of the This means that OMt is most probably not located at the interface
emulsion stabilized with hydrophobic OMt were almost ten times as depicted in the inlay graphs of Fig. 5. Related to this, interfacial
higher than those of the emulsion with hydrophilic NMt and kept tension measurement supports the interfacial rheological results
increasing with the concentration of OMt. The storage modulus obtained for both clay samples [40]. The interfacial tension (␴)
(G’emulsion ) showed a plateau-like behavior over the entire fre- of oil droplet was hardly changed by the addition of OMt even
quency region between 0.1–100 rad/s. With OMt of 3000 ppm, the at increased OMt concentration (for pure oil drop, ␴ is 31 mN/m,
emulsion was not stable and showed phase separation of the water oil droplet with OMt 3000 ppm has 27.2 mN/m, and for that with
phase at the bottom. With higher concentration of OMt (6000 and OMt 5000 ppm, it slightly decreased to 26 mN/m). On the other
20,000 ppm), the emulsions were stable (Fig. 1, Fig. 2(c), and sup- hand, for the oil droplet stabilized with NMt, the interfacial tension
plementary Fig. 2). Since hydrophobic OMt had the unbalanced decreases with increasing concentration of NMt (supplementary
wetting behavior between the oil and the water phase and higher Table 1). According to the previous study [41], OMt accumulated
chemical attraction with the oil phase [39], OMt would remain only above the interface due to high chemical interaction and never
only in the oil phase regardless its concentration. If the amount desorbed to the water phase. Based on both bulk and interfacial
of OMt was recalculated based on the oil phase only (i.e. according rheology measurements, as well as optical observations, the stabi-
to wtclays /wtoil ), the concentration of OMt in the oil phase is quite lization behavior of clays was ensured depending on their surface
significant and enough to increase the modulus of the oil phase properties. The stabilization behavior of OMt rather depended on
(G*oil ) significantly (Fig. 4). If OMt (3000 ppm) added in 40/60 wt/wt the concentration dependent flow properties of the bulk oil phase
O/W emulsion was only dispersed in the oil phase, the concen- (clay dispersed in the oil phase) than on their interfacial rheologi-
tration increases to 7500 ppm (wtclays /wtoil ) for the oil phase. The cal properties. Meanwhile, NMt particles stabilized the emulsion by
J.S. Hong, P. Fischer / Colloids and Surfaces A: Physicochem. Eng. Aspects 508 (2016) 316–326 321

Fig. 6. Oscillatory time sweep test showing interfacial storage and loss modulus (G’interface and G”interface ) of the oil-water interface with the clay mixture (NMt/OMt
3000 ppm/3000 ppm). Pictures of the interface with NMt/OMt (right top) and NMt (right bottom) were taken at 3000 s after the build-up of the interface.

Fig. 7. Oscillatory frequency sweep test showing (a) storage and (b) loss modulus (G’emulsion and G”emulsion ) of emulsions with the clay mixture (NMt/OMt 3000 ppm/3000 ppm)
depending on the composition of emulsion.

the interfacial localization even though only a limited concentra- layer (see picture in the upper right of Fig. 6). They continuously
tion was located at the interface and the interfacial structure was developed an associative structure without desorption [41]. Asso-
not strong. ciative interaction of heterogeneous mixture of clay samples led
to a network with increasing modulus as a function of time. The
3.2. Mixtures of NMt and OMt clays density of networking between clays might increase as clays occu-
pied the interface more. It made the complex interface sustainable
As shown in the previous section, both clay samples did not against external hydrodynamic stress, which resulted in strong sta-
build up a strong interfacial layer to stabilize the emulsion. In order bilization of the emulsion. It means that the interface with high
to generate an interfacial layer for stabilization, a mixture of both interfacial modulus (G*interface ) is strong enough to resist against
clay samples showing different stabilizing effect was tested. The deformation of the contacting phases or to depress the coalescence
combination of both clay samples was fixed to 3000 ppm/3000 ppm induced by net reduction of interfacial energy.
(NMt/OMt) and they were added to the emulsion. Interfacial modu- Bulk rheology of the emulsions stabilized with clay mixture was
lus (G*interface ) of the oil-water interface with the clay mixture was measured depending on the oil-in-water ratio from 30/70 wt/wt
measured as a function of time (Fig. 6). O/W to 70/30 wt/wt O/W (Fig. 7). For ratios up to 50/50 wt/wt O/W,
Different from the interface with pure NMt, the interfacial the storage modulus (G’emulsion ) showed a plateau behavior over
storage modulus (G’interface ) rapidly increased and exceeded the the entire frequency region. The storage modulus (G’emulsion ) of
interfacial loss modulus (G”interface ). It means that the mixture of the 40/60 wt/wt O/W emulsion with the clay mixture increased as
clays rapidly adsorbs at interfacial region and formed an interfacial high as the emulsion with OMt (∼100 Pa) (see also Fig. 3(a)). When
322 J.S. Hong, P. Fischer / Colloids and Surfaces A: Physicochem. Eng. Aspects 508 (2016) 316–326

Fig. 8. Oscillatory frequency sweep test depicting the (a) storage and (b) loss modulus (G’emulsion and G”emulsion ) of the 40/60 wt/wt O/W emulsions with the clay mixture
(NMt/OMt) depending on the composition of two clay samples.

the ratio increased further to 70/30 wt/wt O/W, the modulus of the bromide, CH3 (CH2 )15 N(CH3 )3 Br4 , CAS no.57-09-0, Sigma-Aldrich)
emulsions (G*emulsion ) rapidly decreased below those values of the and SDS (Sodium dodecyl sulfate, C12 H25 NaO4 S, CAS no.151-21-3,
30/70 wt/wt O/W emulsion even though the composition of the Sigma-Aldrich). The frequency sweep experiments of the emul-
viscous oil phase was increased. It means that as the interfacial sions with clay and surfactant were compared in Fig. 9.
area decreases the emulsion separates regardless of the interfacial With the addition of CTAB (3000 ppm) or NMt (10,000 ppm),
modulus (G*interface ). Except for the 40/60 wt/wt O/W emulsion, the the emulsion modulus (G*emulsion ) was lower than 10 Pa and the
emulsions exhibited the phase separation of the water phase at the emulsions were not stable (see Section 3.1). If they were added
bottom or the oil phase at the top (refer to the vial image in Fig. 7 together at the same time, the emulsion modulus (G*emulsion ) had
and supplementary Fig. 4). In other words, the complex interface profoundly increased and the emulsion was significantly stable.
only of the clay mixture (NMt/OMt 3000 ppm/3000 ppm) provides Due to cationic exchange capability of NMt (∼78 meq/100 g dry
strong macroscopic/microscopic stabilization of the emulsion by clay), CTAB enhanced the association of NMt located at the interface
the increased mechanical interfacial property. However, due to the to build up a strong interfacial layer [30], which was sufficient to
lack of electrostatic repulsion, it was shown only at the compo- stabilize the emulsion against coalescence (Fig. 9(d)). On the other
sition of 40/60 wt/wt O/W emulsion (∼45/55 (v/v) O/W) with the hand, if the emulsion was added with OMt and CTAB, the modulus of
high interfacial area, otherwise the emulsions were not preserved the emulsion decreased and the emulsion was not stable (Fig. 9(f)).
from phase separation. The modulus of emulsions (G*emulsion ) sta- Different from the emulsion stabilized with pure OMt (Fig. 3), the
bilized by the complex interface would change only depending on location of OMt in the emulsion is changed by the presence of CTAB
the interfacial modulus (G*interface ), which also showed the max- as the surfactant which manipulates the wetting behavior of OMt
imum value when the droplets coherently flocculated with the between oil and water [30,39]. Coherent location of OMt in the oil
complex interface. If the interfacial layer was formed by the asso- phase to induce the emulsion stabilization could not expect any
ciation between clays, the interfacial modulus (G*interface ) would more and, as a consequence, the emulsion was no longer stable.
change according to the composition of the clay mixture. The When anionic SDS was tested with both clays, the emulsions never
40/60 wt/wt O/W emulsions with different combinations of NMt stabilized. NMt has no interaction with anionic SDS and also the
and OMt was tested (Fig. 8) (the total concentration of clays was rapid adsorption of the SDS to the interface changed the localization
fixed at 6000 ppm). of clay [30]. To achieve a synergistic stabilization effect between
When both clay samples were added equivalently (1/1 clay and surfactant, it is necessary to investigate the surfactant-
NMt/OMt) or when the concentration of OMt was higher than that colloidal clay particles complexation as well as the difference in
of NMt (1/3 NMt/OMt), the emulsion modulus (G*emulsion ) was high adsorption kinetics between them. The strong and rapid interfacial
enough to expect significant stabilization [42]. When the concen- adsorption of surfactant enhances desorption of clays [30], which
tration of OMt was lower than that of NMt (3/1 NMt/OMt), the may result in the phase separation of the emulsion. According to
emulsion modulus (G*emulsion ) was significantly decreased and the the rheological properties of the emulsion, stabilization depends
emulsion showed phase separation. In this case, the interaction on the concentration of clay, their combination, and the presence of
between two clay samples was not strong enough to retain NMt surfactants. Two effects were observed as either clay stabilized the
at the interface and then NMt might be desorbed to the phase hav- emulsion by interfacial layer formation (increase of G*interface ) or by
ing strong chemical affinity with NMt. It means that the complex an increase in the modulus of the oil phase (G*oil ). If the clays had a
interface is sustained only at the limited combination between NMt strong chemical affinity with the oil phase, they did not localize at
and OMt for the emulsion stabilization. the interface but the emulsion modulus (G*emulsion ) increased with
increasing concentrations of clays especially when the phase con-
3.3. Influence of surfactant on emulsion stability taining OMt is the major phase. If clay was localized at the interface,
the emulsion modulus of the emulsion (G*emulsion ) was increased by
In previous work we have shown that the interfacial localization the high interfacial modulus (G*interface ) of the interface. It showed
of clay particles can be influenced by the presence of surfactant and the peak at the optimal concentration of clays since the complex
thus forming a surfactant-clay compound at the interface [30]. To interface was generated by the association between natural clay,
investigate the stabilization behavior of clay with surfactant, both organically modified clay, or surfactants. As the interfacial area of
clay samples were tested with CTAB (Cetyltrimethylammonium
J.S. Hong, P. Fischer / Colloids and Surfaces A: Physicochem. Eng. Aspects 508 (2016) 316–326 323

Fig. 9. (a) Storage modulus (G’emulsion ) and (b) loss modulus (G”emulsion ) as a function of frequency for 40/60 wt/wt O/W emulsion stabilized with clay and surfactant (CTAB
and SDS). Optical inspection of the emulsions after the rheological tests: (c) emulsion with NMt (10,000 ppm), (d) with NMt (10,000 ppm)/CTAB (3000 ppm), (e) with OMt
(10,000 ppm), and (f) with OMt (10,000 ppm)/CTAB (3000 ppm).

and Dinterface are interfacial modulus and interfacial deformation)


as well as each phase (G*water and G*oil for the water (␾water ) and
the oil phase (␾oil ) respectively, here ␾ is the faction of each phase)
[1]:
∗ ∗
 ∗

␶emulsion = Gemulsion D = Goil oil + Gwater water D + ␶interface (1)

Since both bulk phases (oil and water) of the emulsion have
negligible modulus, the modulus of emulsions is determined by
the contribution of interfacial stress (␶interface = G*interface Dinterface )
(␶emulsion ∼ ␶interface ). If the emulsion is not compatibilized, the
interfacial stress might be determined by interfacial energy and
interfacial area (␶interface ∼ interfacial tension (␴) × A/V, where A
and V are interfacial area and volume) [1,28,43]. The emulsion
modulus (G*emulsion ) therefore increases with the increasing com-
position of the dispersed phase because of the increase in interfacial
area. Meanwhile, if particles or macromolecules accumulate at the
interface, the interface has independent interfacial stress regardless
of the contacting phases because it has its interfacial flow proper-
ties (surface dilatational viscosity (␬), surface shear viscosity (␩))
by particle motion, interaction, density as follows [22,44]:
I
  D
 
int = I (2) + f Dint + f Dint (2)

I , and DD are interfacial tension, isotropic sur-


where ␴, I(2) , Dint int
Fig. 10. Comparison of storage and loss modulus (G’emulsion , G”emulsion ) of the emul- face tensor of second order, isotropic and deviatory parts of the
sion (40/60 wt/wt O/W) as a function of clay concentration. Storage modulus of the
rate of strain tensor, respectively. Because the term f(Dint ) is a
oil-water interface with clays (G’interface ) is compared in the table.
function of surface tensor and rate of strain tensor, the emulsion
the complex interface increased, it induced a pronounced stabiliza- modulus (G*emulsion ) will be significantly influenced by interfacial
tion effect to the emulsion. flow properties of its stabilizing adsorption layer. Even if particles
added to the emulsion, they do not necessarily form an interfa-
4. Discussion cial adsorption layer but also disperse only in the oil phase. Thus
particles significantly increase the rheological property of the oil
Rheological response of emulsions (G*emulsion ) under deforma- phase and influence the determination of the emulsion morphology
tion (D) is determined by contribution of the interface (G*interface due to change in viscosity ratio, which might change the emulsion
324 J.S. Hong, P. Fischer / Colloids and Surfaces A: Physicochem. Eng. Aspects 508 (2016) 316–326

Fig. 11. Optical microscope image of (a)(b) the 40/60 wt/wt O/W emulsions and (c)(d) the 60/40 wt/wt O/W emulsions with clay (6000 ppm): (a)(c) (1/1)NMt/OMt and (b)(d)
OMt.

modulus (G*emulsion ) in addition to interfacial stress. Depending at the interface and shown distinct differences in interfacial stor-
on whether the particles locate at the interface, the rheological age modulus as summarized in the table in Fig. 10. Emulsions with
response of the emulsion has totally a different tendency toward NMt or NMt/OMt are therefore easily manipulated by the concen-
the composition of emulsion, the concentration of particles, or tration of particles. The comparison of storage modulus (G’emulsion )
interaction between interfacial localized particles. and loss modulus (G”emulsion ) of the emulsions shown in Fig. 10
According to this study, the O/W emulsions showed different reveals that the emulsion with OMt or NMt has slightly lower stor-
rheological response depending on clay samples. The emulsion age modulus than loss modulus. It means that their structure is
modulus (G*emulsion ) of the 40/60 wt/wt O/W emulsions is com- dissipative under the deformation and the interaction between par-
pared in Fig. 10 as a function of clay concentration and clay types, ticles is not strong enough to resist deformation. The emulsion with
while the morphology of the emulsions is depicted in Fig. 11. NMt/OMt has a higher storage modulus than loss modulus resulting
Because the composition of the emulsion was fixed to 40/60 wt/wt in a strong associative network structure, which does not dissi-
O/W, the interfacial area did not change significantly with the addi- pate under deformation. With OMt, the emulsion shows typical
tion of clay. For OMt, the emulsion modulus (G*emulsion ) steadily spherical structure (Fig. 11(b), (d)) and the emulsion with NMt/OMt
increased with increasing concentration of clays, while the emul- shows non-spherical structure due to rigid association of interfacial
sion with NMt or NMt/OMt shows a peak in emulsion modulus localized clays (Fig. 11(a), (c)).
(G*emulsion ) at 6000 ppm. The difference in the emulsion modulus This rheological study suggests two different stabilization
(G*emulsion ) between OMt and NMt or NMt/OMt is mainly caused mechanisms by the added clay samples. For organically modified
by the location of the clays. OMt disperses well in the oil phase clay (OMt), distinguished rheological properties of the oil phase
and it increases the modulus of the oil phase (G*oil ) with increasing containing clay effectively delay the diffusion of domains and pre-
concentration of clay (Figs. 4 and 5). NMt or NMt/OMt is located vent coalescence when the oil phase is the major phase. On the
J.S. Hong, P. Fischer / Colloids and Surfaces A: Physicochem. Eng. Aspects 508 (2016) 316–326 325

other hand, natural clays (NMt) stabilize the emulsion due to the References
formation of droplets flocculation induced by interfacial localiza-
tion [39]. It is difficult to expect stabilization behavior until clays [1] R.G. Larson, The Structure and Rheology of Complex Fluids, Oxford University
Press, New York, 1999.
coherently occupy the interface. When droplets approach, the com- [2] Z. Zapryanov, A.K. Malhotra, N. Aderangi, D.T. Wasan, Emulsion stability: an
plex interface prevents coalescence and the droplets flocculate to analysis of the effects of bulk and interfacial properties on film mobility and
the cream phase due to their weak colloidal properties. However, drainage rate, Int. J. Multiph. Flow 9 (2) (1983) 105–129.
[3] B.P. Binks, Particles as surfactants-similarities and differences, Adv. Colloid
the interfacial layer by NMt is not strong enough to resist mechan- Int. Sci. 7 (2002) 21–41.
ical stress during emulsification. The addition of both, OMt and [4] B.P. Binks, W. Lie, J.A. Rodrigues, Novel stabilization of emulsions via the
cationic surfactants enhances the interfacial localization, adsorp- heteroaggregation of nanoparticles, Langmuir 24 (2008) 4443–4446.
[5] B.P. Binks, S.O. Lumsdon, Effects of oil type and aqueous phase composition
tion, and aggregation of NMt, which enhances the mechanical on oil-water mixtures containing particles of intermediate hydrophobicity,
strength of the interface. This suggests that natural clays can give Phys. Chem. Chem. Phys. 2 (2000) 2959–2967.
a significant stabilization behavior if they co-exist with substances [6] K.D. Danov, S.D. Stoyanov, N.K. Vitanov, I.B. Ivanov, Role of surfactants on the
approaching velocity of two small emulsion drops, J. Colloid Int. Sci. 368
such as surfactants to induce strong interfacial interaction between
(2012) 342–355.
clays [30,41,42,45–48]. [7] A. Gelot, W. Friesen, H.A. Hamza, Emulsification of oil and water in the
To further understand the stabilization nature of clays, studies presence of finely divided solids and surface-active agents, Colloids Surf. A 12
on the organic compound-clay complex interface have to be carried (1984) 271–303.
[8] A. Tsugita, S. Takemoto, K. Mori, T. Yoneya, Y. Otani, Studies on O/W
out with different clay minerals as well as surfactants. Substantially, emulsions stabilized with insoluble montmorillonite-organic complexes,
in order to obtain a stable emulsion, the emulsion system needs to Colloid Int. Sci. 95 (2) (1983) 551–560.
maintain the structure for the stabilization homogeneously even [9] D.E. Tambe, M.M. Sharma, Factors controlling the stability of
colloid-stabilized emulsions, Colloid Int. Sci. 157 (1993) 244–253.
though each phase plays a critical role in building-up the structure [10] P. Erni, P. Fischer, E.J. Windhab, Sorbitan tristearate layers at the air/water
for stabilization. Accordingly the bulk properties of the emulsion interface studied by shear and dilatational interfacial rheology, Langmuir 21
(G*emulsion ) are significantly useful to know the macroscopic struc- (2005) 10555–10563.
[11] P. Fischer, P. Erni, Emulsion drops in external flow fields − The role of liquid
tural information. The interfacial modulus (G*interface ) also provides interfaces, Curr. Opin. Colloid Interface Sci. 12 (2007) 196–205.
important information about the build-up of the interface, kinet- [12] F. Fenouillot, P. Cassagnau, J.C. Majeste, Uneven distribution of nanaoparticles
ics and interaction between particles. Then, rheology study of the in immiscible fluids: morphology development in polymer blends, Polymer
50 (2009) 1333–1350.
emulsions provides macroscopic and microscopic structural infor- [13] C.P. Whitby, D. Fornasiero, J. Ralston, Effect of oil soluble surfactant in
mation, such as interfacial localization, the role of particles to emulsions stabilized by clay particles, J. Colloid Interface. Sci. 323 (2008)
flocculation, which would be difficult to obtain only from morphol- 410–419.
[14] C.P. Whitby, P.C. Garcia, Time-dependent rheology of clay particle-Stabilized
ogy observations.
emulsions, Appl. Clay Sci. 96 (2014) 56–59.
[15] S. Abend, N. Bonnke, U. Gutschner, G. Lagaly, Stabilization of emulsions by
heterocoagulation of clay minerals and layered double hydroxides, Colloid
5. Conclusions Polym. Sci. 276 (1998) 730–737.
[16] B.P. Binks, S.O. Lumsdon, Catastropic phase inversion of water-in-oil
emulsions stabilized by hydrophobic silica, Langmuir 16 (2000) 2539–2547.
The stabilizing behavior of clay particles in emulsions was stud- [17] S.O. Asekomhe, R. Chiang, J.H. Masliyah, J.A.W. Elliott, Some observations on
ied based on the rheological response of the emulsions depending the contraction behavior of a water-in-oil drop with attached solids, Ind. Eng.
on the localization, concentration, combination of clays, and pres- Chem. Res. 44 (2005) 1241–1249.
[18] V.B. Menon, D.T. Wasan, A review of the factors affecting the stability of
ence of surfactants. Clays stabilized the emulsion by interfacial
solids-Stabilized emulsions, Sep. Sci. Technol. 23 (1988) 2131–2142.
layer formation (G*interface ) or an increase of the bulk oil modu- [19] F.O. Opawale, D.J. Burgess, Influence of interfacial properties of lipophilic
lus (G*oil ). If clay was simply adsorbed at the interface without surfactants on water-in-oil emulsion stability, J. Colloid Interface Sci. 197
strong interaction, the interfacial adsorption layer (G*interface ) was (1998) 142–150.
[20] Y.L. Chen, C.A. Helm, J.N. Israelachvili, Measurements of the elastic properties
not strong enough to prevent coalescence, which also resulted in of surfactants and lipid monolayers, Langmuir 7 (1991) 2694–2699.
low bulk moduli of the emulsion. Formation of a complex inter- [21] D.W. Criddle, A.L. Meader, Viscosity and elasticity of oil surfaces and oil-water
face was obtained by strong association between two clays with interfaces, J. App. Phys. 26 (7) (1955) 838–842.
[22] L.E. Scriven, Dynamics of a fluid interface, Chem. Eng. Sci. 12 (1960) 98–108.
different surface properties or between clay and CTAB, which sig- [23] E. Dickinson, B.S. Murray, G. Stainsby, Coalescence stability of emulsion-sized
nificantly increased the interfacial modulus (G*interface ). Thus, the droplets at a planar oil-Water interface and the relationship to protein film
adsorption layer at the interface was strong enough to provide a surface rheology, J. Chem. Soc. Faraday Trans. I 84 (3) (1988) 871–883.
[24] T. Tadros, Application of rheology for assessment and prediction of the
barrier against droplet coalescence. Further the emulsion modu- long-Term physical stability of emulsions, Adv. Colloid Interface Sci. 108–109
lus (G*emulsion ) increased significantly and stabilized the emulsion. (2004) 227–258.
Both mechanisms give a significant macroscopic/microscopic sta- [25] P. Erni, P. Fischer, E.J. Windhab, V. Kusnezov, H. Stettin, J. Lauger, Stress- and
strain-controlled measurements of interfacial shear viscosity and
bilization to the emulsion as the interfacial area increased which viscoelasticity at liquid/liquid and fas/liquid interfaces, Rev. Sci. Instrum. 74
resulted in high storage modulus of the emulsion (∼100 Pa for (11) (2003) 4916–4924.
G’ of the 40/60 wt/wt O/W emulsion with 6000 ppm OMt or [26] J. Maldonado-Valderrama, J.M. Rodrigues Patino, Interfacial rheology of
protein-surfactant mixtures, Curr. Opin. Colloid Interface Sci. 15 (2010)
3000 ppm/3000 ppm NMt/OMt).
271–282.
[27] H.J. Rivas, P. Sherman, Soy and meat proteins as emulsion stabilizers. 4.The
stability and interfacial rheology of O/W emulsions stabilized by soy and meat
Acknowledgement protein fractions, Colloids Surf. 11 (1984) 155–171.
[28] R. Pal, Rheological behavior of surfactant-flocculated water-in-Oil emulsions:
colloids and surfaces A: physico, Eng. Aspects 71 (1993) 173–185.
This study was supported by Mid-career Researcher Program
[29] P. Mongondry, C.W. Macosko, T. Moaddel, Rheology of highly concentrated
through NRF grant funded by the Korea government (MEST) (No. anionic surfactants, Rheol. Acta 45 (2006) 891–898.
20110016890). [30] J.S. Hong, P.A. Rühs, P. Fischer, Localization of clay particles at the oil-water
interface in the presence of surfactants, Rheol. Acta 54 (2015) 725–734.
[31] Z. Yan, J.A. Elliott, J.H. Masliyah, Roles of various bitumen components in the
stability of water-in-Diluted-Bitumen emulsions, Colloid Int. Sci. 220 (1999)
Appendix A. Supplementary data 329–337.
[32] N. Yan, M.R. Gray, J.H. Masliyah, On water-in-ol emulsions stabilized by fine
Supplementary data associated with this article can be found, in solids, Colloids Suf. A 193 (2001) 97–107.
[33] S. Arditty, V. Schmitt, J. Giermanska-Kahn, F. Leal-Calderon, Materials based
the online version, at http://dx.doi.org/10.1016/j.colsurfa.2016.08.
on solid-stabilized emulsions, Colloid Int. Sci. 275 (2004) 659–664.
040.
326 J.S. Hong, P. Fischer / Colloids and Surfaces A: Physicochem. Eng. Aspects 508 (2016) 316–326

[34] N.P. Ashby, B.P. Binks, Pickering emulsions stabilised by laponite clay [42] T.Y. Jeon, J.S. Hong, Stabilization of O/W emulsion with
particles, Phys. Chem. Chem. Phys. 2 (2000) 5640–5646. hydrophilic/hydrophobic clay particles, Colloid Polym. Sci. 292 (2014)
[35] M.A. Malusis, C.D. Shackelford, H.W. Olsen, Flow and transport through clay 2939–2947.
membrane barriers, Eng. Geol. 70 (2003) 235–248. [43] R. Pal, Rheology of simple and multiple emulsions, Curr. Opin.
[36] R. Suresh, S.N. Borkar, V.A. Sawant, V.S. Shende, S.K. Dimble, Nanoclay drug Colloid.Interface Sci. 16 (2011) 41–60.
delivery system, Int. J. Pharm. Sci. Nanotech. 3 (2) (2010) 901–905. [44] J.V. Boussinesq, Sur l’existence d’une viscosite superficielle, dans la mince
[37] P.C. Schorling, D.G. Kessel, I. Rahimian, Influence of the crude oil couche de transition separant un liquid d’un autre fluide contigu, Ann. Chim.
resin/asphaltene ratio on the stability of oil/water emulsions, Colloids Surf., A Phys. 29 (1913) 349–357.
152 (1999) 95–102. [45] J. Lyklema, Adsorption of ionic surfactants on clay minerals and new insights
[38] A.P. Sullivan, P.K. Kilpatrick, The effects of inorganic solid particles on water in hydrophobic interactions, Prog. Colloid Polym. Sci. 95 (1994) 91–97.
and crude oil emulsion stability, Ind. Eng. Chem. Res. 41 (2002) 3389–3404. [46] H. van Olphen, An Introduction to Clay Colloid Chemistry, John Wiley & Sons,
[39] J.K. Kim, P.A. Rühs, P. Fischer, J.S. Hong, Interfacial localization of nanoclay Inc, USA, 1977.
particles in oil-in-Water emulsions and its reflection in interfacial moduli, [47] F. Bergaya, G. Lagaly, Handbook of Clay Science, Elsevier, Netherlands, 2013,
Rheol. Acta 52 (4) (2013) 327–335. pp. 139–172.
[40] J.S. Hong, J.G. Kim, Variation of interfacial tension by nanoclay particles in [48] H. Katepalli, A. Bose, Response of surfactant stabilized oil-in-water emulsions
oil-in-water emulsions, Compos. Interfaces 21 (8) (2014) 703–713. to the addition of particles in an aqueous suspension, Langmuir 30 (2014)
[41] J.S. Hong, Aggregation of hydrophilic/hydrophobic montmorillonites at 12736–12742.
oil-water interface, App. Clay Sci. 119 (2016) 257–265.

Você também pode gostar