You are on page 1of 20

International Journal of Pharmaceutics 436 (2012) 359–378

Contents lists available at SciVerse ScienceDirect

International Journal of Pharmaceutics


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

Review

Proteins, polysaccharides, and their complexes used as stabilizers for emulsions:


Alternatives to synthetic surfactants in the pharmaceutical field?
Eléonore Bouyer a,b , Ghozlene Mekhloufi a,b , Véronique Rosilio a,b , Jean-Louis Grossiord a,b ,
Florence Agnely a,b,∗
a
Univ Paris Sud, Faculté de Pharmacie, 5 rue Jean-Baptiste Clément, 92296, Châtenay-Malabry Cedex, France
b
CNRS, UMR 8612, Institut Galien Paris-Sud, 5 rue Jean-Baptiste Clément, 92296 Châtenay-Malabry Cedex, France

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

Article history: Emulsions are widely used in pharmaceutics for the encapsulation, solubilization, entrapment, and con-
Received 1 August 2011 trolled delivery of active ingredients. In order to answer the increasing demand for clean label excipients,
Received in revised form 22 June 2012 natural polymers can replace the potentially irritative synthetic surfactants used in emulsion formula-
Accepted 22 June 2012
tion. Indeed, biopolymers are currently used in the food industry to stabilize emulsions, and they appear
Available online 1 July 2012
as promising candidates in the pharmaceutical field too. All proteins and some polysaccharides are able to
adsorb at a globule surface, thus decreasing the interfacial tension and enhancing the interfacial elasticity.
Keywords:
However, most polysaccharides stabilize emulsions simply by increasing the viscosity of the continuous
Emulsion
Stabilization
phase. Proteins and polysaccharides may also be associated either through covalent bonding or electro-
Biopolymer static interactions. The combination of the properties of these biopolymers under appropriate conditions
Drug delivery systems leads to increased emulsion stability. Alternative layers of oppositely charged biopolymers can also be
Layer-by-layer formed around the globules to obtain multi-layered “membranes”. These layers can provide electro-
static and steric stabilization thus improving thermal stability and resistance to external treatment. The
novel biopolymer-stabilized emulsions have a great potential in the pharmaceutical field for encapsula-
tion, controlled digestion, and targeted release although several challenging issues such as storage and
bacteriological concerns still need to be addressed.
© 2012 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
2. Emulsion stability and characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
2.1. Emulsion classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
2.2. Destabilization mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
2.3. Stabilization methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362
2.4. Emulsion characterization and stability assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362
2.4.1. Emulsion type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362
2.4.2. Emulsion granulometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362
2.4.3. Emulsion stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
2.4.4. Rheological properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
2.4.5. Zeta potential measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
2.4.6. Determination of biopolymer adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
2.4.7. Interfacial tension measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364
2.4.8. Interfacial rheology measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364
2.4.9. Interfacial film thickness and refractive index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364

∗ Corresponding author at: Institut Galien Paris-Sud, UMR CNRS 8612, Faculté de Pharmacie, 5 rue Jean-Baptiste Clément, 92296, Châtenay-Malabry Cedex, France.
Tel.: +33 1 46 83 56 26; fax: +33 1 46 83 58 82.
E-mail address: florence.agnely@u-psud.fr (F. Agnely).

0378-5173/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ijpharm.2012.06.052
360 E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378

3. Stabilizing emulsions with biopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364


3.1. Protein-stabilized emulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
3.1.1. Caseins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
3.1.2. Whey proteins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
3.1.3. Gelatin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
3.1.4. Pea proteins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
3.2. Polysaccharide-stabilized emulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
3.2.1. Non-adsorbing polysaccharides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
3.2.2. Adsorbing polysaccharides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
3.3. Protein–polysaccharide stabilized emulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369
3.3.1. Covalent complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369
3.3.2. Non-covalent complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369
3.3.3. Without complexation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369
4. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 370
4.1. Comparison of biopolymers and small-molecule surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 370
4.1.1. Comparison of the interfacial behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 370
4.1.2. Comparison of emulsifying properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372
4.1.3. Comparison of emulsion stabilizing properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372
4.2. Promising applications of biopolymers for the stabilization of pharmaceutical emulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372
4.3. Challenging issues raised by biopolymers in the field of pharmaceutical emulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374

1. Introduction tolerated molecules would thus be a major breakthrough. It is sur-


prising to note that up to now, pharmaceutical emulsions have not
evolved much compared to food ones. In the food industry, the
In the pharmaceutical field, emulsions are widely used as drug introduction of natural polymers such as proteins and polysaccha-
delivery systems for parenteral, oral, and topical (skin and eye) rides as emulsifying and/or stabilizing agents has been a successful
routes (Becher, 1983). They consist of mixtures of at least one approach for the formulation of highly stable emulsions. This was
liquid dispersed in another one in the form of droplets, both liquids allowed thanks to a better understanding of the physicochem-
being immiscible, or poorly miscible one with the other (Tadros istry of emulsion stability, which in turn was made possible by the
and Vincent, 1983). The main interest of emulsions is to encapsu- introduction of new characterization techniques such as interfacial
late a hydrophilic or lipophilic active molecule inside the dispersed rheology.
phase, thus ensuring its protection against environmental stress This review aims at summarizing the contribution of these
and degradation (oxygen, light, enzymes, acidity etc.), and allow- natural stabilizing agents especially in food emulsions, and to dis-
ing its controlled delivery. Emulsions can also mask an unpleasant cuss how this knowledge could be implemented to formulate a new
smell or taste (Ley, 2008), reduce a drug toxicity (Brime et al., 2002), generation of emulsions for pharmaceutical applications that will
enhance its penetration through the skin (Kogan and Garti, 2006) answer the increasing demand for natural excipients and sustain-
or behave as detoxifying systems to entrap toxic molecules (pollu- able development. The first part of this review gives a survey of
tants, pesticides, etc.) in the inner phase (Grossiord and Stambouli, the most recent research on emulsion stability on a physicochem-
2008). In addition, the formation of emulsions can be the inital ical standpoint, and presents the most advanced methods used to
step to obtain nano- and micro-particle systems for drug targeting assess this stability. The second part of this article reviews ongo-
(Grigoriev et al., 2008). ing research on emulsion stabilization by biopolymers (proteins,
The main challenge raised by the use of emulsions is their polysaccharides, and their mixtures). In the last part, a compari-
thermodynamic instability. Indeed, they are prone to destabiliza- son is drawn between small-molecule surfactants and biopolymers,
tion and phase separation (Tadros, 2009; Tadros and Vincent, and the potential pharmaceutical applications of biopolymer emul-
1983). Addition of emulsifiers with interfacial and/or thickening sion stabilizers are described and discussed.
properties allows for the emulsion formation and stabilization.
Pharmaceutical emulsions are mostly stabilized by synthetic sur-
factants (Scherlunda et al., 1998). However, the use of these 2. Emulsion stability and characterization
emulsifiers directly or indirectly raises toxicity and environment
issues (Cserháti et al., 2002; Liwarska-Bizukojc et al., 2005). Syn- 2.1. Emulsion classification
thetic surfactants can be intrinsically toxic or may alter the
distribution and elimination of co-administered drugs (Buggins Emulsions can be classified into two groups: simple emulsions
et al., 2007; Cserháti et al., 2002). Most surfactants induce irri- and multiple emulsions. Two classes of simple emulsions can be
tant skin reactions and may cause toxic symptoms in animals defined, namely oil-in-water (O/W), and water-in-oil (W/O) (Fig. 1).
and humans. For example, the possible binding of anionic sur- Multiple emulsions constitute a more sophisticated system. The
factants to proteins, enzymes, and phospholipid membranes may simplest ones are oil-in-water-in-oil (O/W/O) or water-in-oil-in-
result in biochemical changes such as a modification of protein water (W/O/W) double emulsions (Fig. 1). In the latter for example,
structure and dysfunction of enzymes and phospholipid mem- the emulsion is composed of aqueous droplets, which are dispersed
branes (Cserháti et al., 2002; Effendy and Maibach, 1995). According inside oily drops, these oily drops being themselves dispersed in
to cytotoxicity tests, nonionic surfactants have a lower toxic an external aqueous phase (Grossiord and Sellier, 2001). Emul-
effect than cationic, anionic, and amphoteric ones, the toxicity of sions can also be classified according to their droplet size into three
cationic surfactants being the highest (Effendy and Maibach, 1995; categories: macroemulsions, microemulsions, and nanoemulsions.
Vlachy et al., 2009). Replacing synthetic surfactants by much better Macroemulsions are emulsion systems with droplet sizes ranging
E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378 361

relatively small size of these droplets with regards to the optical


a b wavelengths of the visible spectrum implies that many nanoemul-
sions appear optically transparent, even at large droplet volume
simple emulsions

fraction and for large refractive index contrast (Mason et al., 2006;
Solans et al., 2005). However, nanoemulsions tend to be slightly
turbid if droplet diameter approaches 80 nm (McClements, 2011;
Wooster et al., 2008). Above this size, still in the submicron range,
they appear white due to significant multiple scattering. Contra-
O rily to microemulsions, nanoemulsions are not thermodynamically
W
W O stable.

c d
2.2. Destabilization mechanisms
multiple emulsions

During emulsification, the interfacial area between the con-


tinuous and the dispersed phases is considerably augmented
compared to the interface before dispersion. The interfacial free
W energy is therefore significantly increased. In accordance with
O
the thermodynamic dictum that all systems evolve towards their
O
W global energy minimum state, emulsions rapidly tend to phase
W O
separate in order to minimize the interfacial contact area and
free energy. This instability manifests itself by various mecha-
Fig. 1. Schematic representation of (a) oil-in-water, (b) water-in-oil, (c) water-in- nisms: flocculation, coalescence, creaming or sedimentation, and
oil-in-water, and (d) oil-in-water-in-oil emulsions.
Ostwald ripening (Fig. 2). Flocculation and coalescence are the
major destabilization mechanisms (Walstra, 1987). Flocculation is
from 0.1 to 100 !m. These sizes allow light scattering and give their the result of attractive forces between the droplets and leads to the
white color to these systems. Microemulsions generally contain formation of flocs of dispersed phase. It may occur in biopolymer-
both a surfactant and co-surfactant that induce spontaneous for- stabilized emulsions, either by depletion or bridging mechanisms
mation of the system. Microemulsions are often transparent to the (Fig. 3) (Blijdenstein et al., 2004). Depletion flocculation arises
eye, with low viscosity, and are thermodynamically stable (Friberg when biopolymer concentration exceeds a critical value. Indeed,
and Venable, 1983). This stability is due to their very low interfa- the presence of non-adsorbing or excess biopolymer in the contin-
cial tension (enthalpy), typically 10−1 to 10−2 mN/m (Capek, 2004), uous phase increases the attractive force between the droplets by
and to the low enough droplet size (entropy). Indeed, their droplet osmotic effect and provokes the exclusion of the biopolymer chains
size varies from 100 Å to 100 nm (Solans et al., 1997). The last from the narrow region surrounding two droplets (McClements,
category, nanoemulsions, refers to emulsions with droplet sizes in 2000). This depletion mechanism is not specific to biopolymers.
the nanometric scale, i.e. with a mean diameter of 20–200 nm. The Indeed, it can also be induced by non-adsorbed small surfactant

coalescence

flocculation

O
W

coalescence
and
phase
separation

Fig. 2. Schematic representation of the various destabilization mechanisms for an oil-in-water emulsion.
362 E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378

a b environment as previously discussed. Finally, the fourth class com-


prises polymer surfactants. In addition to their general interfacial
tension lowering properties, polymers induce steric or electro-
static interactions, changes in the interface viscosity or elasticity,
O O or changes in the bulk viscosity of the system thus improving emul-
sion stability.
Biopolymers (proteins and polysaccharides) belong to this
W W last class. They prevent flocculation and coalescence by com-
bined mechanisms. Biopolymer gels and non-adsorbing thickening
polysaccharides reduce droplets movements and encounters in
emulsions by increasing the viscosity of the continuous phase.
Adsorbing biopolymers act on inter-droplet interactions. Indeed,
Fig. 3. Schematic representation of (a) depletion and (b) bridging flocculation in
flocculation and coalescence can be avoided by inducing repulsive
presence of polymer.
electrostatic interactions or steric hindrance between droplets.
Emulsion stabilization or destabilization by a biopolymer
micelles, and by inorganic or organic nanoparticules. Bridging floc- depends on various parameters, such as polymer nature, poly-
culation occurs when a single biopolymer molecule adsorbs at mer concentration, pH, ionic strength, etc. When formulating
the surface of more than one emulsion droplet. It thus acts as a biopolymer-stabilized emulsion, it is thus necessary to con-
a polymeric link and promotes bridging flocculation (Guzey and trol them. The emulsifying process is also a key parameter that
McClements, 2006a). The formed flocs are the first step to droplet influences emulsion microstructure (droplet size). Emulsions can
sedimentation (in W/O emulsions) or creaming (in O/W emulsions), be formed by various protocols, involving rotor-stator systems,
and coalescence. ultrasound, high-pressure systems, phase inversion temperature,
In some cases, the flocculation mechanism is reversible and micropores, etc. (Schubert and Engel, 2004). However, it is out of the
simple magnetic stirring or hand shaking can break the flocs. scope of this review to give a detailed description of these processes.
However, in some other cases, it is irreversible and droplets coa-
lesce. Small droplets fuse together and progressively form larger 2.4. Emulsion characterization and stability assessment
globules. Emulsions can also undergo creaming or sedimentation,
which consist in the migration of globules to the top or bottom This section gives a general overview of the most common and
of the dispersion, depending on the relative densities of the two advanced techniques used for the characterization of emulsions
liquid phases. Finally, Ostwald ripening is a diffusion mechanism and for the elucidation of their stabilization mechanisms. However,
of dispersed phase molecules through the continuous phase, with- the following list is not exhaustive.
out any contact between droplets, related to the solubility of the
dispersed phase in the external phase. This diffusion takes place 2.4.1. Emulsion type
from small droplets towards larger ones, driven by a higher Laplace The continuous phase of an emulsion can be water (in O/W
pressure existing in the former. Over time, the droplet size distribu- emulsion) or oil (in W/O emulsion). However, the general aspect
tion shifts toward larger values. Ultimately, these mechanisms can of both systems is often very similar, even if some sensorial tech-
lead to the oiling off and the emulsion breaks up into separated niques rapidly allow the identification of one or the other emulsion
oil and water phases. Flocculation, sedimentation and cream- type (shining aspect, oily touch, dilution in water, dissolution of
ing are reversible mechanisms of droplets migration, whereas methylene blue) (Groeneweg et al., 1993). Measurement of the
Ostwald ripening and coalescence are non-reversible mechanisms external phase conductivity is more accurate. Indeed, an aque-
of droplets size increase. For an in-depth understanding of these ous continuous phase has a higher conductivity than an oily one
destabilization mechanisms, a very good description can be found (Suttiprasit et al., 1993). Most biopolymers are water soluble, and
in the review authored by Damodaran (2005). thus tend to form O/W emulsions.
In the pharmaceutical field, it is necessary to control these
destabilization mechanisms. Depending on emulsions application, 2.4.2. Emulsion granulometry
it may be important to slow them down or to determine the exact The granulometry of an emulsion is a very important parameter,
environmental conditions (shear rate, pH, etc.) allowing emulsion which is related to emulsion process. The droplet size distribu-
destabilization and consequent release of the encapsulated active tion and polydispersity can be measured by means of various
pharmaceutical ingredient at the right time, at the targeted site. experimental techniques, such as laser diffraction (Gharsallaoui
et al., 2010a; Li et al., 2010), dynamic light scattering (Guzey and
2.3. Stabilization methods McClements, 2006b; Ogawa et al., 2004), and optical microscopy
(Guzey and McClements, 2006b). The most popular definitions
Emulsifiers and stabilizers are commonly used to kinetically available to obtain the droplet-size are (Brittain, 2001):
stabilize emulsions. They are classified in four categories (Myers, !
1999). The first class is composed of non-surfactant ions that can ni di3
adsorb at the droplet surface without affecting the interfacial ten- the volume − surface(or Sauter) mean diameter : d32 =
i
!
sion or facilitating the emulsification process. However, some may, ni di2
under appropriate conditions, impose a slight electrostatic barrier i
between approaching droplets thus contributing to emulsion sta- (1)
bilization. The second class is composed of small non-surfactant
colloidal solids (i.e. silica or clay particles of Pickering emulsions) !
that adsorb and form a physical barrier between droplets, thereby ni di4
delaying or preventing coalescence. The third class is constituted
i
of classical monomeric surfactants such as sodium dodecyl sul- the volume − moment mean diameter : d43 = ! (2)
phate. They decrease interfacial tension and increase emulsion ni di3
stability but can be quite irritative and potentially toxic towards the i
E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378 363

where ni and di are the number and diameter of the globule popu- stability of emulsions, their texture, syringeability, and condition-
lation i, respectively; ing possibilities for future applications (Tadros, 1994).
kB T
and the hydrodynamic radius : Rh = (3) 2.4.5. Zeta potential measurements
6!"D
Zeta potential measurements may help predict flocculation and
where kB is the Boltzmann constant, T the temperature, " the vis- coalescence because emulsion stability can be influenced by the
cosity of the continuous phase, and D the diffusion coefficient of a interfacial electric charges (Gharsallaoui et al., 2010a; Guzey and
droplet. McClements, 2006b; Li et al., 2010; Ogawa et al., 2004). The effec-
tive charge of a droplet differs from its actual charge because of
2.4.3. Emulsion stability the presence of counter-ions in the solution. The closest counter-
Emulsion stability is a result of physical stability (no phase ions are attached by electrostatic attractions. They are in immediate
separation), chemical stability (no chemical reaction), and micro- contact with the droplet surface and form the Stern layer. They thus
biological stability (no germ development). The physical stability decrease the droplet actual charge. Their concentration then rapidly
can be evaluated by two accelerated aging methods, which mimic decreases when deviating from the droplet surfaces till reaching
conditions that emulsions could undergo during transport and stor- their equilibrium solution concentration. In an electrical field, the
age. The first one consists in evaluating samples stored at different adsorbed counter-ions of the Stern layer move jointly with the
temperatures (from 0 to 40 ◦ C) and for various periods of time (from droplet. The electrophoretic mobility ($ep ) is first calculated using
24 h to 15 days). It may also include repeated hot–cold cycles. The the von Smoluchowski theory modified by Henry (Hunter, 1986)
second method enhances by centrifugation the sedimentation or (Eq. 5), with v, the droplets rate and E, the applied electrical field.
creaming processes that can lead to coalescence (Tcholakova et al.,
v
2005). Accelerated aging studies by centrifugation allow determi- $ep = (5)
E
nation of the emulsion stability index (ES) following Eq. (4):
The zeta (%) potential can then be calculated applying Henry’s
remaining emulsion height relation (Eq. 6):
ES(%) = × 100 (4)
initial emulsion height
2ε%f (ka)
$ep = (6)
However, one must be careful when drawing conclusions from 3"
results obtained by these methods, as they do not perfectly mimic
the natural aging process. where ε is the liquid dielectric constant, " its viscosity, and f(ka)
Another relatively new approach is to evaluate emulsion sta- is Henry’s function being equal to 1.5 when approximating with
bility by using a Turbiscan (Formulaction® , L’Union, France). This Smoluchowski. The %-potential corresponds to the potential mea-
apparatus is composed of a reading head consisting of a pulsed near sured at the Stern layer. A high absolute value of the %-potential
infrared light source (# = 850 nm) and two synchronous detectors avoids flocculation. Conversely, a small value (close to zero) allows
that scan the entire length of an emulsion filled tube. The trans- the droplets to come in closer contact with each other and floccu-
mission detector measures the light flux transmitted (at 0◦ angle) late.
through the emulsion, and the backscattering detector receives The ionic strength is of importance in emulsion stability as high
the light backscattered (at 135◦ angle) by the emulsion. Diffuse ionic strength can induce droplet flocculation if the electrolyte con-
reflectance and transmittance versus sample height and time are centration is high enough to electrostatically screen droplet charge
obtained, allowing detection of flocculation and coalescence pro- (Laplante et al., 2005). This screening can easily be detected by
cesses even at an early stage (invisible to the eye) (Lemarchand %-potential measurement because its value gets closer to zero.
et al., 2003; Mengual et al., 1999).
2.4.6. Determination of biopolymer adsorption
2.4.4. Rheological properties It is possible to plot adsorption isotherms of a biopolymer at the
The flow characteristics of emulsions are influenced by various oil–water interface. The graph represents the amount of adsorbed
interaction forces that occur in the system during sedimenta- polymer versus its bulk concentration (Graham and Phillips, 1979;
tion or creaming, flocculation, coalescence, and Ostwald ripening. Jenkins and Ralston, 1998). The shape of the curve indicates when
Several factors affect emulsion rheology such as volume fraction the interface is entirely covered by ending as a plateau. It also gives
and viscosity of the dispersed phase, droplet granulometry, viscos- information about the adsorbed molecule conformation and about
ity of the continuous phase as well as its chemical composition competitive adsorption (Chatelier and Minton, 1996).
(polarity, pH), nature and electrolyte concentration, and inter- In the case of proteins, their interfacial concentration (' ) can
facial rheology (Tadros, 1994). Numerous methods are available also be calculated from the specific surface area of the oil droplets,
to investigate emulsion flow properties. It is possible to perform and from the difference between the amount of protein used to pre-
flow, creep, and oscillatory experiments. In oscillatory measure- pare the emulsion and that measured in the aqueous phase after
ments, a viscoelastic material is subjected to a sinusoidally varying the emulsification process. After a centrifugation step, the follow-
strain (or stress). The response in stress (respectively strain) is then ing mass balance (Eq. 7) is frequently used (Euston et al., 2000;
monitored. From these measurements, it is possible to determine Tcholakova et al., 2005, 2006; Ye, 2010):
emulsion viscoelastic parameters such as storage (G% ) and loss mod- VC (CINI − CSER ) (1 − ˚)d32
uli (G%% ). G% modulus represents the elastically stored energy in the ' = = (CINI − CSER ) (7)
SVOIL 6˚
emulsion, and G%% modulus the energy dissipated in viscous flow
(Tadros, 1994). Depending on their properties, emulsions can either where S is the specific surface area of the droplets, VC and VOIL are
maintain the necessary fluidity for their transport and distribu- volumes of the continuous and oil phases, respectively, ˚ is the oil
tion or exhibit a yield behavior inhibiting flow under small shear volume fraction, CINI is the initial protein concentration in the solu-
stress. The shear-thinning behavior is probably the most encoun- tion, and CSER is the protein concentration in the aqueous phase (in
tered rheological behavior in the field of emulsions: the apparent the serum) after emulsification determined by the Bradford method
viscosity decreases when increasing the shear rate. From the varia- (Bradford, 1976). It is important to note that to perform such a
tion of the different rheological parameters, it is possible to infer the characterization, the equilibrium between the adsorbed and the
364 E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378

non-adsorbed chains should not be modified by the centrifugation are very well stabilized because of the formation of viscoelastic
step. adsorbed layers (Bos and Vliet, 2001; McClements, 2004).

2.4.7. Interfacial tension measurements 2.4.9. Interfacial film thickness and refractive index
Three techniques can allow interfacial tension measurements: The thickness of an interfacial film can be determined by
the drop method, based on the analysis of the shape of a drop, the ellipsometry. This technique can be applied to the adsorption of
Wilhelmy plate and the Du Noüy ring methods consisting in mea- surfactants molecules at the air-liquid interface (Binks et al., 2000;
suring the force that can pull out a plate or ring from the studied Castellani et al., 2010) as well as the oil–water interface (Binks
liquid surface. The pendant (or rising) drop and Wilhelmy plate et al., 2003; Bylaite et al., 2001). The principle of the measure-
techniques allow the determination of the adsorption kinetics of ments is based on the change in the state of polarization of the
an amphiphilic molecule at the interface, which is of importance reflected polarized light against the studied interface. This change
for the emulsification process (Ward and Tordai, 1946; Graham and is expressed in amplitude ratio (( ) and phase shift ()). These two
Phillips, 1979; Demé et al., 1995; Rosilio et al., 1997). parameters give the refractive index of the two materials provided
that the interface is clean. In the case of an adsorbed interfacial film,
2.4.8. Interfacial rheology measurements they depend on its thickness and refractive index. It is then possible
The interfacial rheology technique is based on surface modifi- to determine the adsorbed layer thickness, and its mean refractive
cation induced by mechanical forces, such as shear or dilational index, knowing these parameters (( , )) for several surrounding
stresses, giving the functional relationship between the deforma- media. The monitoring of ( and ) allows studying the process of
tion of the interface (gas/liquid or liquid/liquid interfaces), the interfacial layer formation upon adsorption of an emulsifying agent
stresses applied in and on it, and the resulting flows in the adja- (Bylaite et al., 2001; Russev et al., 2000). For example, Russev et al.
cent fluid phases (Ivanov et al., 2005; Krägel and Derkatch, 2010; (2000) obtained a thickness of 1.8 nm for the first "-casein layer
Ravera et al., 2010). It is suitable in a wide range of technical adsorbed at the oil–water interface, and 5.4 nm for the second one.
applications involving multiphase systems with a high specific It is thus possible to apply this technique to a multi-layered inter-
area such as emulsions and foams. This technique is relatively facial film. It is also possible to show evidence of the destabilization
new and still not developed enough. Few or no studies of emul- of an interface, for example upon dilution (Henni et al., 2005).
sification and emulsion stability, wetting or two-phase flow have
been compared to interfacial shear rheological properties. How-
3. Stabilizing emulsions with biopolymers
ever, interfacial rheological techniques give very useful information
on inter- and intramolecular interactions at interfaces (Krägel and
Biopolymers such as proteins and polysaccharides are good
Derkatch, 2010).
examples of natural emulsifiers/stabilizers currently encountered
Stress/strain relationships occurring at interfaces of bi-
in the food industry and could be used to stabilize pharma-
dimensional phases such as surfactants or biopolymers single-
ceutical emulsions. In the pharmaceutical field, these natural
/multi-layers in an emulsion can be studied by interfacial rheology.
polymers are already used for other applications such as cap-
Elastic and viscous moduli are obtained for each interfacial strain.
sule formation (gelatin), tablet binder (gum arabic, chitosan,
The drop method can be used for dilational rheological measure-
hypromellose), suspending agent (gum arabic, hypromellose),
ments. The volume of a pendant (or a rising) drop is submitted to
mucoadhesive formulations (chitosan), matrix for extended-
sinusoidal variations, thus resulting in expanding or compressing
release tablets (hypromellose), etc. (Rowe et al., 2006). Many
the surface of the drop. The interfacial tension is determined by
proteins can act as emulsifiers because of their ability to adsorb
analyzing the profile of the drop according to the Laplace equation,
at the oil–water interface (Möbius and Miller, 1998) and to
by means of a Charge Coupled Device camera connected to a com-
increase emulsion stability (McClements, 2004). Most polysac-
puter. Another technique consists in using a rheometer equipped
charides behave as emulsion stabilizers by forming an extended
with a bi-conical disk. This disk is placed at the interface between
network in the continuous phase which thus becomes highly vis-
the two liquids and rotates to determine the viscous and elastic
cous (Dickinson, 2009) and can even form a gel (Benna-Zayani
shear rheological parameters of the studied interface. A complex
et al., 2008; Weiss et al., 2005). Only few polysaccharide deriva-
modulus (E*) can be defined for both techniques as follows (Eq. 8):
tives possess surface properties that enable their adsorption at the
E ∗ = E % + iE %% (8) oil–water interface (Dickinson, 2003; Garti and Reichman, 1993).
Combining the properties of proteins and polysaccharides under
where E% is the elastic modulus describing the elastic properties of appropriate conditions (concentration, protein-to-polysaccharide
the interface, and E%% the viscous modulus correlated to its viscous ratio, pH, ionic strength, and temperature) may be a valuable
resistance. The phase angle (ı) is related to E% and E%% moduli by Eq. strategy for improving emulsion stability (Dickinson, 1995,
(9): 2008a,b; Guzey and McClements, 2006a). Indeed, proteins and
polysaccharides can form complexes through covalent binding or
E %%
tan ı = (9) attractive electrostatic interactions. In the latter case, alternative
E%
layers of oppositely charged biopolymers can be formed around
where ı values below 45◦ reveal a predominantly elastic behavior, a the oil globules to obtain multi-layered “membranes”. These layers
more favourable condition for emulsion stability. Correspondingly, provide both electrostatic and steric stabilizations thus improving
a high elastic modulus compared to the viscous one is correlated thermal stability and resistance to external treatment (i.e. ther-
to good emulsion stability. Bos and Vliet (2001) consider that dila- mal processing, chilling, and freezing) (Gu et al., 2007; Guzey and
tional elasticity and viscosity are very important indicators of film McClements, 2006a; Thanasukarn et al., 2006).
rupture, since this rupture can mainly be seen as a dilational defor- The most recent studies on emulsion formulation and stabiliza-
mation. If we consider emulsion stability, this means that when two tion by proteins, polysaccharides, and their mixtures in the food
droplets of the emulsion encounter, the elastic interface resists to area are presented below. These biopolymers are classified accord-
the strain stress created by the deformation of the droplets, thus ing to their main emulsifying/stabilizing mechanism, and to the
preventing coalescence and maintaining emulsion stability (Laza- nature of the interactions in the case of protein–polysaccharide
Knoerr et al., 2011). Emulsions containing adsorbing biopolymers mixtures. The examples of biopolymers presented below are
E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378 365

mainly those already used in pharmaceutical applications other et al., 1994). Casein-stabilized emulsions do, however, destabi-
than emulsion formulation. Their use in pharmaceutical emulsions lize by a flocculation mechanism in presence of calcium salts, as
should therefore be more easily transferable. observed by Dickinson and Davies (1999) by droplet-size mea-
surements. Dalgleish and others (Agboola and Dalgleish, 1995;
3.1. Protein-stabilized emulsions Dalgleish, 1997) explained this phenomenon by the strong spe-
cific calcium ion binding to the negatively charged phosphoserine
Many proteins are able to adsorb at the interface and facili- groups of casein, which reduces the net negative charge and thick-
tate droplets disruption by lowering the interfacial tension (Norde, ness of the adsorbed caseins layer. To our knowledge, no marketed
2003). However, at equivalent interfacial concentration small syn- pharmaceutical emulsion containing caseins exists to this date, but
thetic surfactants are generally more effective than proteins in these proteins can be used in some nutritional supplementation
reducing the interfacial tension. Typically, most proteins decrease formulations.
the oil–water tension by about 15–20 mN/m at saturated mono-
layer coverage compared to 30–40 mN/m for small synthetic 3.1.2. Whey proteins
surfactants (Razumovsky and Damodaran, 1999). However, pro- Milk globular whey proteins are a mixture of #-lactalbumin, "-
teins lead to thermodynamically and kinetically more stable lactoglobulin, and several other minor proteins (immunoglobulins,
emulsions (Bos and Vliet, 2001; Damodaran, 2005). In their primary serum albumin) that are less flexible than caseins. They are
structure, proteins contain hydrophilic and hydrophobic residues currently used with success in food emulsions, especially "-
randomly spread all over the structure. In the tertiary folded con- lactoglobulin (18,300 g/mol) and #-lactalbumin (14,200 g/mol)
formation, some of these residues exist as segregated patches. At (Dickinson, 1997; Tcholakova et al., 2006). For example, Kim et al.
the oil–water or air–water interface, protein adsorption proceeds (2002) have shown that it was possible to stabilize emulsions
in three main stages (Graham and Phillips, 1979): diffusion from with "-lactoglobulin at pH 7, as this protein bears a net nega-
the bulk to the vicinity of the interface, actual adsorption, and tive charge and induces relatively strong electrostatic repulsion
reorganization of the adsorbed protein. During adsorption, only a between droplets. Indeed, the pI for this protein is 5.2. However,
fraction of the protein hydrophobic groups is embedded into the these authors also observed by particle size measurements that "-
lipophilic phase and most of the protein structure remains into lactoglobulin-stabilized emulsions were rather heat sensitive and
the aqueous one. Proteins then undergo conformational changes flocculated during heat treatments (Kim et al., 2002). This finding
in order to maximize the number of favorable interactions and is in agreement with Ye’s results (2010). Temperature should not
minimize the number of unfavorable ones in their environment. exceed a certain critical value (∼70 ◦ C) otherwise proteins unfold
Intermolecular interactions between adsorbed protein molecules and expose reactive groups that increase attractive interactions
can lead to the formation of a highly viscoelastic film that prevents between them (Kim et al., 2002), leading to extensive droplet
coalescence (Dickinson, 1999). The remaining protein residues in flocculation (McClements, 2004). Beside its stabilizing effect by
the aqueous phase also provide steric stabilization against floccu- electrostatical interactions, "-lactoglobulin is also able to decrease
lation and coalescence (Wilde et al., 2004). It is important to note the air water or oil–water interfacial tension (Paulsson and Dejmek,
that all proteins possess various amounts of hydrophobic groups 1992; Wüstnek et al., 1999). Moreover, Wüstnek et al. (1999) have
and exhibit different flexibility. Consequently, they exert diverse performed interfacial dilational rheology measurements using the
interfacial properties. pendant drop method and reported an elastic behavior for this pro-
Typically, 1–3 wt% protein concentration is used to stabilize tein. They described that at low protein concentrations, when the
emulsions (Damodaran, 2005). In this concentration range, protein first biopolymer molecules reach the interface in optimal condi-
adsorption rate is not a limiting factor to emulsion formation. The tions, they can fully adsorb at any locus and occupy a maximum
efficiency of proteins in emulsion stabilization can be improved interface area, because the interface is still free of molecules. How-
through physical, enzymatic, and genetic modifications. For exam- ever, at higher concentrations, abundant lateral protein–protein
ple, partial denaturation or unfolding of proteins can be obtained interactions at the interface induce a more compact conformation
under controlled heating and shearing conditions (Demitriades of the protein. At interfacial saturation, elastic and viscous moduli
et al., 1997; Mitidieri and Wagner, 2002). These modifications allow reach a maximum and for higher interfacial concentrations, both
them to expose more hydrophobic groups towards the air or oil moduli decrease. In a previous work, we also observed this behav-
phase. ior in the oil–"-lactoglobulin solution system (Bouyer et al., 2011).
Among others, caseins, whey proteins, gelatins, and pea proteins Bovine serum albumin (66,338 g/mol (Peter, 1975)) can also form
are commonly used in the food industry. We describe them in more interfacial elastic films, as inferred from pendant drop and sur-
details in the following sections. face pressure-area measurements (Burgess and Ozlen Sahin, 1997;
Serrien et al., 1992; Ward and Regan, 1980).
3.1.1. Caseins To this day, there is no example of "-lactoglobulin and #-
Caseins are the main protein components of mammalian milk. In lactalbumin applications in the pharmaceutical industry. Human
emulsion formulation, sodium caseinate is generally used. It mainly serum albumin, however, can be used in hypovolemia to increase
consists of a partially aggregated mixture of four individual flexi- intravascular volume by increasing plasma colloid oncotic pres-
ble caseins (#s1 , #s2 , ", $) (19,000–25,000 g/mol) (Dickinson and sure. The increased intravascular pressure is the consequence of
Davies, 1999). In milk, caseins are intimately associated with cal- increased protein concentration in the intravascular spaces upon
cium phosphate in the form of colloidal particles called “casein albumin infusion (Copelin, 1998).
micelles” (Dickinson, 2006). The two major individual caseins, #s1
and ", are extremely effective emulsifying agents. They are able 3.1.3. Gelatin
to decrease the interfacial tension during emulsification, and to Gelatin is a relatively high molecular weight protein derived
protect newly formed fine droplets against flocculation and coa- from animal collagen, e.g. pig, cow or fish (Ledward, 1986).
lescence by a combination of electrostatic and steric repulsions The latter is often preferred because there is no concern about
(Dickinson and Davies, 1999). Moreover, adsorbed "-casein lay- contamination with bovine spongiform encephalitis, and because
ers at the air–water interface display a viscoelastic behavior as ethnic groups that do not consume pig or cow can eat it. Collagen
evidenced by magnetic rod interfacial shear rheometry (Bantchev is hydrolyzed by boiling in the presence of acid (Type A gelatin,
and Schwartz, 2003) and electrocapillary wave diffraction (Gau pI ∼ 7–9) or alkaline (Type B gelatin, pI ∼ 5) (Taherian et al., 2011).
366 E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378

Gelatin is slightly surface active and can act as an emulsifier in oil- (Alderman, 1984). Casas et al. (2000) have reported that depending
in-water emulsions (Lobo, 2002). Surh et al. (2006) have confirmed on fermentation conditions, the content in acetate and pyruvate
that physically stable oil in water emulsions could be obtained with groups varies, as well as the average molecular weight of the gum
fish gelatin at concentrations equal to or above 4.0 wt%. However, (from 1.5 × 106 to 5 × 106 g/mol), influencing the viscosity of its
these authors observed a small fraction of relatively large droplets solutions. In water, the stiff polymer chain may exist as a single,
(>10 !m). They explained it by the relatively low surface activity double or triple helix that interacts with another chain to form
of fish gelatin compared to "-lactoglobulin, for example. Gelatin is a complex, loosely bound network (Jansson et al., 1975; Melton
currently used as a binder in emulsions containing light sensitive et al., 1976). This particular structure gives the gum its unusual
silver halide salts for photographic films (Lourenço and Sampaio, thickening properties, with a yield stress and shear-thinning and
2009; Sheppard, 1929). In the pharmaceutical field, gelatin remains thixotropic behaviors (Benmouffok-Benbelkacem et al., 2010). Due
mostly used in the manufacture of capsules (Jones, 2004). to these unique rheological properties, xanthan gum is recognized
as an excellent emulsion stabilizer (Hennock et al., 1984; Sun et al.,
3.1.4. Pea proteins 2007). Under appropriate concentration conditions, the gum is used
The two major protein components of pea proteins are the to prevent flocculation and creaming in emulsions. Several authors
storage globulins: 7S (vicilin, 150,000–180,000 g/mol) and 11S have calculated the minimum yield stress value (* 0 ) that an emul-
(legumin, 360,000 g/mol) (Gueguen et al., 1988). They have a regu- sion should have at a given droplet radius in order to inhibit gravity
lar quaternary structure, trimeric for vicilin proteins and hexameric effects such as creaming. A general way of estimating * 0 is reported
for legumin ones. It has been shown that pea proteins are able in Eq. (10) (Benna-Zayani et al., 2008; Chhabra, 1993; Jossic and
to decrease the oil–water interfacial tension and contribute to Magnin, 2005).
emulsions stabilization by the formation of a rigid membrane at
*0 ≥ Y)+gR (10)
the interface (Ducel et al., 2004a,b). However, pea proteins sta-
bilized emulsions can be very sensitive to environmental factors with Y a dimensionless number describing the yield stress
such as pH and ionic strength. For this reason, they are usually effects to gravity (6 × 10−2 to 7 × 10−2 ), )+ the density difference
combined with another emulsifier/stabilizer such as high methoxyl between the two phases, g the gravity acceleration, and R the mean
pectin (Gharsallaoui et al., 2010a,b). To the authors’ knowledge, radius of the droplets. Knowing the mean radius allows for the
no marketed pharmaceutical product containing pea proteins determination of the necessary minimum yield stress value for sta-
exists. bilizing emulsions. When xanthan gum is mixed with guar gum or
magnesium aluminium, some synergistic rheological effects on the
3.2. Polysaccharide-stabilized emulsions viscosity occur (increased viscosity or gelation) (Kovacs, 1973), as
well as when mixed with locust bean gum (Benna-Zayani et al.,
Polysaccharides are known for their water-holding and thick- 2008).
ening properties because of their hydrophilic character and high In the pharmaceutical field, xanthan gum is used as a suspend-
molecular weight. They can be classified in two categories for their ing and stabilizing agent in oral and topical formulations (Singh,
use in stabilizing emulsions droplets. Most common polysaccha- 2006a), for the production of sustained-release matrix tablets (Billa
rides do not have much of a tendency to adsorb at fluid interfaces. et al., 2000; Ceulemans et al., 2002; El-Gazayerly, 2003; Peh and
Non-adsorbing polysaccharides have no or limited surface activity Wong, 2000; Phaechamud and Ritthidej, 2007; Yeole et al., 2006)
and enhance the emulsion stability by gelling or modifying the vis- or for its mucoadhesive properties (Gacesa, 1988).
cosity of the aqueous continuous phase, which slows down droplet
movement and encounters (Paraskevopoulou et al., 2004). Some 3.2.1.2. Alginates. Alginates are hydrophilic colloidal carbohy-
other polysaccharides, such as gum arabic, chemically modified drates extracted with dilute alkali from various species of
starch or cellulose derivatives, naturally occurring galactoman- brown seaweeds (Phaeophyceae), especially Laminaria hyper-
nan hydrocolloids (guar gum, fenugreek gum), acetylated pectin borean, Macrocystis pyrifera, and Ascophyllum nodosum. The
from sugar beet, etc. (Dickinson, 2003) display surface/interfacial chemical composition of alginates is directly dependent on the
activity. They first stabilize emulsions by adsorption at the oil biological source, growth, and environmental conditions. The struc-
droplet surface and then prevent droplet flocculation and coales- ture of alginate is mainly composed by alternating blocks of
cence through electrostatic and/or steric repulsive forces. For gum #-(1-4)-l-guluronic acid (G), "-(1-4)-d-mannuronic acid (M), and
arabic and galactomannans, the surface activity results mostly from mixed GM blocks (Gacesa, 1988). The pKa values of G and M
the presence of a protein fraction in their structure. It also seems monomers have been reported to be 3.38 and 3.65, respectively, in
that the protein associated with the pectin plays an important 0.1 M NaCl (Haug, 1964). Several authors showed that the stiffness
role in stabilizing emulsion. Conversely, for cellulose derivatives, of the three blocks decreased in the order GG > MM > GM (Mackie
the surface activity is due to the combination of hydrophobic and et al., 1980; Smidsrød, 1970). Sodium alginate can give highly vis-
hydrophilic groups along the cellulose backbone. In the follow- cous solutions. It has unique colloidal properties, which include
ing subsections, polysaccharides are classified in non-adsorbing thickening, stabilizing, suspending, film forming, gel producing,
polysaccharides if they do not exhibit (or only limited) surface and emulsion stabilizing (Fabra et al., 2008; Moe et al., 1995). Diva-
activity and adsorbing ones if they can stabilize emulsions through lent ions (such as Ca2+ ) form ionic bridges with glucuronic acid
their adsorption at the oil–water interface by lowering the interfa- units belonging to different chains (Gomez-Diaz and Navata, 2004).
cial tension. This kind of ionic association, so-called “the egg-box model”, leads
to the formation of physical gels (Morris et al., 1978). Hambleton
3.2.1. Non-adsorbing polysaccharides et al. (2009a) have reported the very promising results obtained
3.2.1.1. Xanthan gum. Xanthan gum is an anionic polysaccharide with alginate emulsion-based edible films to protect microencap-
produced by the bacterium Xanthomonas campestris. Its structure sulated aroma compounds. In the marketed pharmaceutical field,
is composed of a "-(1-4)-d-glucose main chain, and side chains alginates are mainly used as thickeners and gelifiers, in capsules or
each one out of two glucose residues. Side chains are constituted tablets.
of a #-d-mannose, "-d-glucuronic acid, and "-d-mannose as ter-
minal residues. Pyruvic acid and acetate groups can also be found 3.2.1.3. Carrageenans. Carrageenans are key natural gelling
on terminal mannose residues and non-terminal ones, respectively polysaccharides extracted from the cell wall and intracellular
E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378 367

matrix of numerous species of seaweeds (Eucheuma, Chondrus, as protein, peptide, and nucleotide therapeutics (Oh et al., 2010).
and Gigartina). They are a family of sulfated linear polysaccharides Hyaluronic acid is often used in emulsions as the first step of hydro-
(Therkelsen, 1993). There are three major types of carrageenans: gel particles or microspheres synthesis. Drug delivery systems can
kappa ($), iota (%), and lambda (&)-carrageenans. Their primary then be obtained with these hydrogel particles (Dulong et al., 2004;
structure is composed of alternating disaccharide repeating Iglin et al., 2010; Yun et al., 2004). Recently, a hyaluronic acid gel
sequences of "-d-galactose linked at position 3, and #-d-galactose that contained vasoactive intestinal peptide-loaded liposomes was
linked at position 4. They differ in the number/position of sulphate developed (Lajavardi et al., 2009). Interactions between hyaluronic
groups and the content of 3,6-anhydrogalactosyl rings per disac- acid chains and liposomes increased the viscosity of the gel and
charide (Glickman, 1982). Carrageenans are able to form gels on enhanced its elasticity. This resulted in a delayed release of the
cooling. The gelation process starts by the transition of disordered vasoactive intestinal peptide for the treatment of uveitis.
coils to coaxial double helices (McKinnon et al., 1969). Helix–helix
aggregation promoted by cations (such as K+ or Ca2+ ) gives rise to 3.2.1.5. Chitosan. Chitosan is a naturally occurring polysaccharide,
thermal hysteresis between sol–gel and gel–sol transitions (Morris commonly obtained from the alkali or enzymatic deacetylation of
et al., 1980). Kappa and iota are the main carrageenans able to chitin. Chitin is found in the cuticles of crustacean, insects, molluscs,
form gels due to their anhydride bridges allowing conversion of and cell walls of microorganisms. Chitosan structure is composed
coils in double helix structures. Conversely, &-carrageenan has of glucosamine and N-acetylglucosamine copolymers. It is posi-
no anhydride bridge and is therefore non-gelling (Stanley, 1990). tively charged in acidic aqueous solution due to the protonation
It is possible to insert an anhydride bridge by alkali modification of its amino residues, which have a pKa value of 6.2–7.0 (Claesson
(Ciancia et al., 1993) and obtain a theta carrageenan. However, the and Ninham, 1992). Various molecular weights can be found:
latter cannot form a gel either, due to sulfation at O(2) of 3-linked (50–2000) × 103 g/mol (Baldrick, 2010). Chitosan is widely investi-
residues preventing the formation of the double-helix structure. gated for many different applications, including drug delivery sys-
Doyle et al. (2010) have recently confirmed the very low viscosity tems, due to its biodegradability, biocompatibility, bioadhesivity,
of theta carrageenan solutions. antimicrobial activity, and film forming properties (No et al., 2007;
Carrageenans can be used to influence the rheology and skin Shahidi and Abuzaytoun, 2005; Varum and Smidsrød, 2006). Its
permeation of microemulsion formulations. Valenta and Schultz cationic nature enables chitosan to strongly interact with the
(2004) showed that the presence of carrageenans in the con- oppositely charged lipids of biomembranes. It is an amphiphilic
tinueous phase of microemulsions increased the porcine skin polyelectrolyte that combines both electrosteric and viscosifying
permeation of sodium fluorescein by improving the required con- stabilization mechanisms in emulsions (Rodriguez et al., 2002).
sistency for a better application on large skin areas. Like alginates, Chitosan is a very good viscosity-enhancing agent in an acidic envi-
$- and %-carrageenans can be used to encapsulate aroma com- ronment, owing to its high molecular weight and linear unbranched
pounds (Hambleton et al., 2009b; Marcuzzo et al., 2010). They are structure. It has a shear thinning behavior (Singla and Chawla,
often studied in aqueous solutions in combination with another 2001). Skaugrud (1991) reported that the viscosity of chitosan solu-
biopolymer such as "-lactoglobulin (Ould Eleya and Turgeon, tions increased with its concentration and degree of deacetylation,
2000), pea proteins (Musampa et al., 2007), or locust bean gum and decreased with the temperature. Payet and Terentjev (2008)
(Dunstan et al., 2001). They are also combined with other biopoly- showed that an aqueous chitosan solution was able to emulsify
mers for emulsion stabilization (with "-lactoglobulin for example) paraffin oil. They observed chitosan adsorption at the air–water
(Dickinson and Pawlowsky, 1998; Gu et al., 2005, 2007). Stable and oil–water interfaces with limited interfacial tension decrease,
emulsions can be obtained with only 0.1–0.5 w/v % carrageenan and formation of a dense network of polyelectrolytic brushes in
concentrations (Singh, 2006b). Carrageenans are used in many the continuous phase. Chitosan chains reduced the oil droplet dif-
nonparenteral dosage forms such as suspensions, emulsions, gels, fusion by forming a rigid elastic network in the aqueous phase
lotions, eye drops, suppositories, tablets, and capsules (Singh, thus improving emulsion stability. Its low surface properties com-
2006b). pared to its viscosifying ones made us classify chitosan in the
non-adsorbing polysaccharide category. Chitosan can be used in
3.2.1.4. Hyaluronan. Hyaluronan is an anionic linear biopolymer controlled drug delivery systems (Jones and Mawhinney, 2006;
composed of alternating disaccharide units of d-glucuronic acid Sezer and Akbuğa, 1995). Vasconcellos et al. (2011) for example,
and N-acetyl-d-glucosamine with "(1-3) and "(1-4) linkages successfully produced papain-loaded chitosan microparticles for
(Ambrosio et al., 1999). It is ubiquitous in the human body controlled release applications. Chitosan was also proposed as a
(extracellular matrix, synovial fluid, umbilical cord) with excellent component of mucoadhesive dosage forms and in gene delivery
physicochemical and biological properties including biodegrad- systems (Hejazi and Amiji, 2003). It allowed, on its own or in com-
ability, biocompatiblity, non-toxicity, non-immunogenicity, spe- bination with a surfactant, the stabilization of simple and multiple
cific viscoelasticity, hydration, and lubrification (Garg and Hales, emulsions (Chuah et al., 2009; Hino et al., 2000; Moschakis et al.,
2004; Kuo, 2006). Its molecular weight ranges from 1 × 103 to 2010; Schulz et al., 1998). However, chitosan is not in any marketed
1 × 107 g/mol (Oh et al., 2010). In solution, hyaluronan has a drugs because its safety as a pharmaceutical excipient has not been
stiffened random coil structure with little tendency to form cross- assessed yet (Baldrick, 2010). Nevertheless, chitosan is currently
linked gels (Oh et al., 2010) and displays a shear thinning behavior used in dietary preparations for obesity, and has been recognized
(Pisárčik et al., 1995). It has been reported that many proteins can as safe in the US for use in foods (Baldrick, 2010).
bind to hyaluronan (hyaladherins) (Day and Prestwich, 2002), influ-
encing its conformation (Day and Sheehan, 2001). Hyaluronan is 3.2.2. Adsorbing polysaccharides
used in osteoarthritis (by intra-articular injections, known as vis- 3.2.2.1. Gum arabic. Gum arabic is the most well known of all
cosupplementation), ocular surgery (as, for example, a vitreous polysaccharide emulsifiers in food industry. It is widely used
replacement), plastic surgery (generally used as a matrix which in soft drinks for its emulsifying properties (Chanamai and
is injected to treat facial wrinkles and folds, and to enhance the McClements, 2002; Dickinson, 2003; Garti, 1999a; McNamee et al.,
appearance of the lips), tissue engineering (for tissue regeneration, 1998). It is exudated from the Acacia senegal tree (Randall et al.,
restoration, and repair, mainly because of its high biocompatibil- 1989). Gum arabic is a hetero-polysaccharide containing both
ity), etc. (Oh et al., 2010). It is also increasingly studied for targeted protein and polysaccharide subunits. It is composed of three
specific and long-acting delivery of various biopharmaceutics such fractions depending on the relative protein-to-polysaccharide
368 E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378

content: the arabinogalactan (AG) (80–90 wt% of the total gum), form trains which are separated by hydrophilic loops and tails
the glycoprotein (GP) (2–4 wt% of the total gum), and the that extend into the aqueous phase (Wollenweber et al., 2000).
arabinogalactan-protein (AGP) (10–20 wt% of the total gum) frac- Nahringbauer (1995) proposed that, similarly to proteins, adsorp-
tions (Anderson et al., 1983). The molar mass of the entire gum tion was followed by changes in molecular conformation. Initially,
can vary from 3.0 × 105 to 5.8 × 105 g/mol depending on its origin in bulk solution, the macromolecules are coiled and after adsorp-
and tree age (Idris et al., 1998). Aqueous solutions of gum arabic tion the polymer backbone starts to unfold. Pharmaceutical use of
have a rather low viscosity (100 mPa s for a 30% w/v solution at HPMC (0.45–1 w/w %) has been reported in vehicles for eye drops
20 ◦ C (Kibbe, 2006a)) compared to those of other polysaccharides and artificial tear solutions (Harwood, 2006). In oral pharmaceu-
having similar molecular masses. Its rheological behavior is consid- tical products, it is primarily used as a tablet binder (Chowhan,
ered as Newtonian up to 20–30 wt% gum concentrations (Sanchez 1980), film-coating (Banker et al., 1981; Rowe, 1980), or matrix for
et al., 2002). The AGP fraction is considered to be responsible for the extended-release tablets (Dahl et al., 1990; Hogan, 1989). In addi-
emulsifying properties of gum arabic (Erni et al., 2007; Fauconnier tion, HPMC is an alternative to gelatin in the manufacturing of hard
et al., 2000; Picton et al., 2000; Randall et al., 1988). The association capsules (Harwood, 2006).
of polysaccharidic blocks along the peptidic chain gives a “wat-
tle blossom”-type structure and confers to the macromolecule an 3.2.2.3. Galactomannans. Galactomannans are abundant plant
amphiphilic character favoring its adsorption at the air–water and polysaccharides that occur as energy reserves in certain legumi-
oil–water interfaces (Dickinson, 2003; Idris et al., 1998). Indeed, the nous seeds. They are rather rigid hydrophilic biopolymers with
proteinaceous fractions of the gum would embed in the air or oil a polymannose backbone ("-(1-4)-d-mannose) and grafted #-
phase, while the carbohydrate component would extend out from (1-6)-d-galactose units. The mannose to galactose (M/G) ratio
the surface into the aqueous phase (Chanamai and McClements, is dependent on the galactomannan source and environmental
2002; Picton et al., 2000). In emulsion systems, the adsorption of growth conditions. The M/G ratios of fenugreek gum, guar gum,
the gum at the oil–water interface allows steric hindrance pre- tara gum, and locust bean gum are of about 1, 2, 3, and 4, respec-
venting the droplets to come in closer contact. Moreover, the gum tively (Dea and Morrison, 1975; Wu et al., 2009). They can form
being negatively charged above pH 2 (Kravtchenko, 1997) allows highly viscous solutions even at low concentrations (for instance,
for electrostatic stabilization of emulsions. Although its surface a viscosity of 4.86 Pa s for a 1 w/v % guar gum dispersion) (Kibbe,
activity is considered as rather low when compared to typical pro- 2006b). Galactomannans are thus generally used as thickening,
tein emulsifiers (McNamee et al., 1998), we recently showed that water-holding, and stabilizing agents. The rheological behavior of
after sufficient adsorption time, gum arabic lowered the interfacial these solutions is usually non-Newtonian. Their shear-thinning
tension to the same extent as "–lactoglobulin (Bouyer et al., 2011). behavior has been described as following the Cross model (Eq. 11)
In order to generate stable sub-micron-sized droplets, it is neces- (Cross, 1965; Sittikijyothin et al., 2005).
sary to use a rather high gum-to-oil weight ratio, i.e. approximately " #
1:1 (McNamee et al., 1998). "0 − "∞
* = "∞ + p ,̇ (11)
Gum arabic is used in various delivery systems. Recently, Avadi 1 + (ˇ,̇)
et al. (2010) obtained promising results in the development of
insulin-loaded nanoparticles based on ionic gelation of chitosan with * the shear stress, ,̇ the shear rate, "0 and "∞ the zero and
and gum arabic for oral delivery. In another study, gum arabic infinite shear viscosities, ˇ and p the time and rate constants.
was used to encapsulate and control the release of endoglucanase Several groups (Garti, 1999b; Garti and Reichman, 1994;
(Ramakrishnan et al., 2007). Lu et al. (2003) also used the gum as Mikkonen et al., 2009; Wu et al., 2009) have proposed some of
an osmotic, suspending, and expanding agent in an osmotic tablet these galactomannans for emulsion stabilization. Their interfacial
system. It is reported that optimal concentrations of gum arabic tension lowering properties are assumed to be most probably due
are of 10–20 wt% when used as an emulsifying agent, 10–30 wt% as to the presence of a small fraction of protein closely associated with
a pastille base, 5–10 wt% as a suspending agent, and 1–5 wt% as a the polysaccharide structure, as in gum arabic (Brummer et al.,
tablet binder (Kibbe, 2006a). 2003; Dickinson, 2003; Youssef et al., 2009). However, Garti and
Reichman (1994) showed that the protein fraction of galactoman-
3.2.2.2. Hydroxypropylmethylcellulose (HPMC). HPMC, also named nans played no significant role in the gum activity. They suggested
Hypromellose, is part of an interesting group of surface active that the strong interfacial activity of galactomannans was due to
cellulose derivatives which have a strong tendency to accumu- a “salting-out” effect, which helped precipitation of a gum film
late at the oil–water interface causing a reduction of the surface onto the oil droplets. This contradiction between results obtained
tension (Nahringbauer, 1995). It is a partly O-methylated and O- by the different groups is due to the method used for separat-
(2-hydroxypropylated) cellulose. HPMC finds various applications ing the protein fraction from the whole galactomannan molecule.
in pharmaceutical and food industries where it is used for con- Indeed, Garti’s method was a physical separation (ultracentrifu-
trolled drug release, control of texture and rheological properties gation), whereas Brummer et al. (2003) used enzyme hydrolysis
of dispersions, emulsification, etc. The use of HPMC as an emulsifier to remove the proteins. Brummer et al. (2003) showed that the
for oil in water emulsions was first described by Daniels and Barta purified gum was less surface active than the unpurified one. The
(1993). The emulsion stabilizing properties of cellulose derivatives proteins closely associated with galactomannans could not be eas-
could be associated with their structural features. Indeed, the sur- ily separated by physical methods, but were susceptible to enzyme
face activity of these macromolecules is due to their hydrophobic hydrolysis.
(rich in methoxylgroups) and hydrophilic (rich in hydroxypropyl In the pharmaceutical field, galactomannans are binders and
groups) regions distributed along the cellulose backbone, allow- disintegrants in solid-dosage forms. Coviello et al. (2007), for exam-
ing for their adsorption at fluid interfaces and the lowering of ple, formed a tablet matrix from guar and locust bean gums, and
the interfacial tension (Camino et al., 2009; Ochoa-Machiste and studied the release of model molecules. They showed that it was
Buckton, 1996; Wollenweber et al., 2000). HPMCs exhibit differ- possible to modulate the release of the molecules from this galac-
ent surface activities with varying the methoxyl/hydroxypropyl tomannan matrix.
ratio. When adsorbed, the conformation of this biopolymer may
be described according to the train-loop-tail model: hydropho- 3.2.2.4. Pectin. Pectin is a complex polysaccharide extracted from
bic segments along the polymer chain attach to the interface and plant cell walls, especially citrus peels, apple pomace, and sugar
E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378 369

beet pulps. Its structure is mainly composed of esterified d- at 45 ◦ C when emulsions were stabilized by a sodium caseinate-
galacturonic acid residues in an #-(1-4) chain, and depends on maltodextrin conjugate, compared to sodium caseinate alone.
the pectin source and extraction process. This backbone is then The large size of the bound polysaccharide generates long-range
regularly branched with rhamnogalacturonan segments. Pectin is steric repulsion between emulsion droplets surfaces. The differ-
mainly used for its gelling and dispersion stabilizing properties. ent experimental studies of protein–polysaccharide conjugates
In addition, it has emulsifying properties, and as early as 1927, published over the 2004–2007 period have been reviewed by
was proposed for the preparation of vegetable oils emulsions and Dickinson (2009).
mayonnaise (Rooker, 1927). Pectin is able to produce fine and stable
emulsions (Leroux et al., 2003) in the same manner as gum arabic 3.3.2. Non-covalent complexes
but at much lower dosage (Leroux et al., 2003; Siew and Williams, More often protein–polysaccharide complexation arises from
2008). Since it contains protein residues, the model of its associa- non-covalent association, mainly driven by attractive electrostatic
tion to oil droplets may be similar to that of gum arabic (Dickinson, interactions. Numerous studies have shown improved emulsion
2003; Leroux et al., 2003). Thus, pectin emulsification properties stability, attributable to the presence of associative interfacial inter-
have been mainly attributed to its protein content, but also to actions between the protein and polysaccharide. The main studies
its acetyl groups (Leroux et al., 2003; Siew and Williams, 2008). based on this stabilization mechanism are gathered in Table 1. Two
Indeed, the presence of acetyl groups in pectins (4–5%) enhances alternative procedures for emulsion formation can be used. The first
the hydrophobic nature of the polysaccharide. However, the pro- one consists in preparing a mixed solution of the biopolymers, and
tein associated with pectin plays the key role, and the presence of using the resulting protein–polysaccharide complex for the emul-
acetyl groups is not an absolute requirement with respect to the sification (Laplante et al., 2006; Lutz et al., 2009). The second one,
emulsifying efficiency (Leroux et al., 2003). According to Siew and named the layer-by-layer (LBL) electrostatic deposition technique,
Williams (2008), apart from protein content, other factors, such as consists in forming a primary protein-stabilized emulsion and then
the molecular weight, the ferulic acid, acetic acid, and methyl ester in adding the polysaccharide. Complexation occurs directly at the
contents, and their accessibility to the surface of the oil droplets, droplet surface through attractive electrostatic interactions. This
contribute to the emulsification properties of pectins. Pharmaceu- leads to the formation of a second layer and allows surface charge
tical applications of pectin include formulations for the treatment reversal (Khalloufi et al., 2009a). This last step can be repeated
of diarrhea, constipation, and obesity (Cook, 2006). Pectin has using oppositely charged biopolymers to increase the number of
also been studied in gel formulations for oral sustained release of layers coating the oil droplets. It seems important to start by a
ambroxol (Kubo et al., 2004), and in gel beads for controlled drug protein layer because of its faster adsorption resulting in a rapid
delivery within the gastrointestinal tract (Murata et al., 2004). lowering of the interfacial tension (Klein et al., 2010). Under cer-
tain physicochemical conditions, this multilayer coating displays a
3.3. Protein–polysaccharide stabilized emulsions better stability to environmental stresses such as pH, ionic strength,
heating, chilling, freezing, and dehydration, than a single interfa-
The combination in the same system of the advantages of pro- cial layer (Aoki et al., 2005; Gu et al., 2007; Guzey and McClements,
teins (fast adsorption compared to polysaccharides) and polysac- 2006b). The detailed principles of this LBL method can be found
charides (steric repulsion or viscosity enhancement) to stabilize in the review of Guzey and McClements (2006a). The LBL depo-
emulsions is increasingly studied. Indeed, proteins and polysaccha- sition of biopolymers onto the oil droplets surface can be easily
rides both contribute by their emulsifying/stabilizing properties to adapted to high-scale production conditions because of the fast-
create novel emulsions with improved stability and functionality. ness, simplicity, and relative cheapness of this method (Grigoriev
The development of surface active protein–polysaccharide com- and Miller, 2009). However, in order to achieve stable and homo-
plexes can be achieved either by covalent bonding or electrostatic geneous multilayered stabilized emulsions, it is necessary to fully
interactions. In the latter case, it is important to wisely choose control the composition of the system and the conditions of prepa-
the polymers and physicochemical conditions so that the protein ration (protein:polysaccharide ratio, pH, ionic strength, biopolymer
and polysaccharide carry opposite charges. Another way to stabi- concentration, etc.). For example, the presence of free biopolymer
lize emulsions is to combine protein and polysaccharide properties molecules in the aqueous phase during addition of the oppositely
without developing any attractive interactions between the two charged coating biopolymer may lead to the formation of com-
biopolymers. plexes in the solution bulk. These complexes often provide bridging
flocculation and strongly reduce emulsion stability (Surh et al.,
3.3.1. Covalent complexes 2005). They can also lead to depletion flocculation, due to an excess
An extreme type of protein–polysaccharide interaction occurs in polysaccharide. Emulsion stability also depends on the protein
when a covalent linkage is formed between the two biopoly- interfacial coverage: if protein molecules do not saturate the inter-
mers, creating a new amphiphilic biopolymer with improved face, the added polysaccharide can anchor at the globule surface.
surface properties. Protein–polysaccharide conjugates can be This co-adsorption modulates the interfacial stability.
formed without using any chemicals, generally by linking pro-
tein '-amino-groups to the reducing end carbonyl groups of 3.3.3. Without complexation
polysaccharides through controlled Maillard reaction (slow dry This last stabilization mechanism consists in adsorbing first a
heating) (Kato et al., 1990). The newly formed conjugate exhibits biopolymer (generally a protein), and then adding a polysaccharide
substantially improved solubility and emulsifying properties that does not interact with the adsorbed biopolymer but increases
compared to the protein (i.e. "-lactoglobulin + dextran (Dickinson the viscosity of the continuous aqueous phase. The protein allows
and Galaska, 1991; Wooster and Ann Augustin, 2006), whey the formation of small droplets during the emulsification process,
protein isolate + maltodextrin (Akhtar and Dickinson, 2007), and the polysaccharide generates an extended thickening network,
sodium caseinate + maltodextrin (O’Regan and Mulvihill, 2010)) which induces high viscosity at low shear rate, thus slowing down
and polysaccharide alone. Furthermore, the adsorbed protein layer droplet motion. The improved stability can also be explained by
seems to be protected against destabilization under unfavorable a limited thermodynamic compatibility between the two biopoly-
environmental conditions (i.e. heating, freezing, high electrolyte mers, resulting in enhanced interfacial protein adsorption. Many
concentrations, etc.). O’Regan and Mulvihill (2010), for example, examples of such protein–polysaccharide associations can be
have reported improved emulsion stability after seven-day storage found in the literature, among which sodium caseinate + xanthan
370 E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378

Table 1
Main published experimental studies on protein–polysaccharide electrostatic combination to stabilize emulsions.

Protein1 Protein2 or Protein3 or Experimental conditionsa (pH; References


polysaccharide1 polysaccharide2 concentrations; pr:ps ratio; incorporation
order)

Bilayer stabilization Whey protein isolate Flaxseed gum pH 3.5; 0.6 w/v % pr, 0–0.33 w/v % ps; pr1 Khalloufi et al. (2009a)

then ps1
pH 3.5, 7; 0.6 w/v % pr, 0–0.33 w/v % ps; pr1 Khalloufi et al. (2009b)
then ps1
Chitosan – pH 5.5, 6; 0.1% ps; 0:1, 2.5:1, 5:1, 7.5:1, Laplante et al. (2006)
10:1 ratios; pr1 and ps1 premix
Pectin – pH 2–8; 4:0.5 w/w % solution; pr1 and ps1 Lutz et al. (2009)
premix
Xanthan gum – pH 7; 0.5–6 wt% pr, 0.5 wt% ps; pr1 then ps1 Benichou et al. (2007a)
Galactomannans – pH 7; 5 wt% pr; 4:0.5 (wt/wt) ratio; pr1 Benichou et al. (2007a)
then ps1
$-carrageenan – pH 4.5–7.5; 0.5, 3 w/w % pr, 0–0.4 w/w % Singh et al. (2003)
ps; pr1 and ps1 premix
Gum arabic – pH 7; 4:1, 3:1, 2:1, 1:1 (1–10 wt%) ratios; Klein et al. (2010)
pr1 and ps1 premix and pr1 then ps1
Gelatin – pH 3.4, 6.8; 1 wt% pr; 1:1; pr1 then pr2 Ledward (1986)

"-lactoglobulin Pectin – pH 7 → 4; 0.225 wt% pr, 0.2 wt% ps; pr1 Guzey and McClements (2006b)
then ps1
pH 3, 7; 0.225 wt% pr, 0.2 wt% ps; pr1 then Guzey et al. (2004)
ps1
pH 3; 0.45 wt% pr, 0–0.22 wt% ps; pr1 then Moreau et al. (2003)
ps1
Carrageenans – pH 3.5, 6; 0.5 wt% pr, 0–0.15 wt% ps; pr1 Gu et al. (2005)
then ps1
%-carrageenan – pH 3.4, 4, 6; 3:7–3:2 (0.5 wt%) ratios; pr1 Ru et al. (2009)
then ps1
Alginate – pH 3–7; 0.45 wt%; 0–0.5 wt%; pr1 then ps1 Pongsawatmanit et al. (2006)
Gum arabic – pH 4.2; 0.5, 2.5 w/w %; 2:1, 1:2; pr1 and ps1 Bouyer et al. (2011)
premix, pr1 then ps1 , ps1 then pr1
Sodium caseinate Pectin – pH 7; 0.5% pr; 0–0.2% ps; pr1 then ps1 Bonnet et al. (2005)
Gellan gum – pH 5.4; 1% pr; 0.03, 0.05% ps; pr1 and ps1 Sosa-Herrera et al. (2008)
premix
Dextran sulfate – pH 6; 0.5 wt% pr, 0.1–1 wt% ps; pr1 and ps1 Jourdain et al. (2008)
premix and pr1 then ps1
Pea protein Pectin – pH 2.4; 0.25 wt% pr, 0.2 wt% ps; pr1 then Gharsallaoui et al. (2010a)
ps1
Trilayer stabilization "-lactoglobulin %-carrageenan Gelatin pH 6; 0.5 w/w % pr1 , 0.1 w/w % ps1 , 0.2 Gu et al. (2007)
w/w % pr2 ; pr1 then ps1 then pr2
Pectin Chitosan pH 4; 0.1125% pr1 , 0.1% ps1 , 0.15% ps2 ; pr1 Guzey and McClements (2006b)
then ps1 then ps2
Chitosan Pectin pH 3–7; 0.05–0.1 wt% pr1 , 0–0.1 wt% ps1 , Li et al. (2010)
0–0.25 wt% ps2 ; pr1 then ps1 then ps2
Chitosan Alginate pH 3–7; 0.05–0.1 wt% pr1 , 0–0.1 wt% ps1 , Li et al. (2010)
0–0.25 wt% ps2 ; pr1 then ps1 then ps2
a
pr: protein; ps: polysaccharide; pr and ps premix: pr and ps were mixed in solution prior to emulsion formulation; pr then ps: ps was added after emulsion formulation
with pr; ps then pr: pr was added after emulsion formulation with ps. 1 stands for the first pr or ps, 2 stands for the second one, and 3 for the third one.

gum (Hemar et al., 2001; Moschakis et al., 2005), whey protein their structural complexity, and the cooperative nature of their
isolate + xanthan gum (Sun and Gunasekaran, 2009), whey pro- adsorption, adsorption of biopolymers is a much slower process:
tein isolate + flaxseed gum (Khalloufi et al., 2008), and sodium the steady state is reached only after hours or days – or even not at
caseinate + locust bean gum (Perrechil and Cunha, 2010). all (Dickinson, 2011). Whereas the adsorption of small surfactants
is reversible, the adsorption of biopolymers is usually considered
4. Discussion as practically irreversible (from a kinetic and not a thermodynamic
standpoint) due to a very low desorption rate (Bos and Vliet, 2001).
4.1. Comparison of biopolymers and small-molecule surfactants Theoretical desorption rates of proteins are 104 –108 times slower
than for usual surfactants. As a consequence, experiments per-
The aim of this section is to compare the interfacial behavior formed on timescales of 104 s are unable to discriminate between
of small-molecule surfactants and biopolymers and to discuss the reversibility and irreversibility of adsorbed biopolymers (Ferri
implications of their differences on emulsion formation and stabi- et al., 2010). The adsorption of biopolymers also involves consid-
lization. erable changes in their three-dimensional structure, unlike low
molecular weight surfactants for which such structural changes
4.1.1. Comparison of the interfacial behavior and conformational rearrangements are very limited. Indeed, pro-
Because of their smaller size and lower molecular weight, teins and adsorbing polysaccharides do not display clearly defined
surfactants diffuse, adsorb, and stabilize an interface more hydrophilic head and hydrophobic tail as found in low molecular
rapidly than biopolymers. With small-molecule surfactants weight surfactant structure. Thus, the structural and dynamic
true thermodynamic equilibrium between bulk and interface is properties of the adsorbed layers at liquid interface are mostly
reached within seconds. Conversely, owing to their higher size, determined by the degree of unfolding of the biopolymer before
E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378 371

and after adsorption. Proteins and polysaccharides can adsorb at The thickness of the interfacial layer of monomeric amphiphilic
the interface with several segments (trains), and then change their molecules does not vary to a great extent with surface coverage and
conformation and orientation to allow more segments to adsorb remains close to the size of a surfactant molecule. Conversely for
over time. Owing to their higher molecular weight, polysaccharides biopolymers, this parameter strongly depends on the surface cover-
usually adsorb more slowly at the interface than proteins (Bouyer age (Camino et al., 2009; Miller et al., 2000; Perez et al., 2006). At low
et al., 2011). Although the adsorption process is different for low biopolymer interfacial concentration, the protein or the polysac-
molecular weight surfactants and biopolymers, it is interesting to charide can adopt a flat and extended conformation at the interface.
note that the same kinetic approach is used in the literature to When the biopolymer concentration is increased, the molecular
describe the behavior of the two kinds of amphiphilic molecules. At area occupied by one adsorbed molecule decreases and the biopoly-
low surface pressures (low bulk concentration and short adsorp- mer conformation is more condensed. The number and length of the
tion times), diffusion is the rate-determining step and a modified biopolymer loops and tails protruding into the aqueous phase thus
Ward and Tordai equation can be used (Ward and Tordai, 1946): increase, leading to larger values of the layer thickness. The extent
of thickness variations depends on the nature of biopolymer (num-
$ Dt %1/2
ber of effective hydrophobic groups and distance between them,
˘ = ,(t) − ,0 = C0 kB T (12)
! flexibility of the macromolecule backbone, molecular weight, etc.).
At the highest biopolymer bulk concentrations, a collapsed struc-
where ˘ is the surface pressure, ,(t) and , 0 are the surface tension
ture with multilayer formation can be observed (Bouyer et al., 2011;
at time t and at time t = 0 respectively, C0 is the bulk concentration
Bos and Vliet, 2001; Camino et al., 2009).
of the amphiphilic molecule in the aqueous phase, kB is the
The resulting mechanical properties of the interfacial films are
Boltzmann constant, T the absolute temperature, D the diffusion
also very different. The interfacial shear viscosities of biopolymers
coefficient of the surface active molecule, and t the adsorption
are much higher than those of small-molecule surfactants. This
time. If diffusion controls the adsorption process the plot of
was attributed to lateral interactions between biopolymers due
surface pressure versus t1/2 is linear, and the diffusion coefficient
to hydrogen bonding, covalent bonding, hydrophobic and electro-
can be calculated from its slope (Demé et al., 1995). At higher
static interactions (Bos and Vliet, 2001).
interfacial pressures (high concentrations or semi-dilute regime),
Many studies have attempted to combine the different inter-
solute–solute interactions occur and the plot is no longer linear.
facial mechanisms and properties of low molecular weight
To analyse the adsorption, unfolding and rearrangements of
surfactants and biopolymers in order to improve emulsion sta-
molecules occurring at higher surface pressures, the approach pro-
bilization. However, the interfacial mechanisms involved in such
posed by Graham and Phillips (1979) can be used, where the rates
mixed systems are far more complex than for single molecules.
of the different proccesses are analysed by first-order equations of
The competitive displacement of proteins from the interface
the following form:
by small-molecule surfactants has been studied and computer
˘f − ˘t simulated (Courthaudon et al., 1991; Dickinson et al., 1990; Mackie
ln = −ki t (13) et al., 2000; Seta et al., 2012). At monolayer saturation coverage,
˘f − ˘0
only about one-third of the accessible surface area is usually occu-
In this equation, ˘ f , ˘ 0 , and ˘ t are the surface pressures at pied by train segments of a protein. Thus, the numerous small gaps
the end of the adsorption process, at time t = 0, and at time t, in the adsorbed biopolymer layer are readily accessible to small
respectively, and ki the first-order rate constant. When experi- molecules such as surfactants (Dickinson and Matsumara, 1994).
mental data are plotted according to Eq. (13), two or more linear The observed mechanisms depend on the surfactant type, i.e. ionic
regions can be observed. The first one corresponds to the adsorp- or non-ionic, and on its concentration. Non-ionic surfactants ini-
tion, with the first-order constant of adsorption kads . The slope tially adsorb into the void spaces at the interface, which raises the
of the second linear region is taken as the first-order constant of local surface pressure. As the nucleated domains grow with time,
rearrangement (kr ), occuring among a more or less constant num- the surrounding protein film is compressed, and the local surface
ber of adsorbed molecules. For low molecular weight surfactants pressure increases more. At sufficiently high surface pressures (i.e.
this rearrangement process is negligible. The two above described high surfactant concentrations), the protein film looses its integrity
models have been recently applied to the determination of the and the protein desorbs in the bulk phase. Complete removal
kinetics of adsorption of various biopolymers: a mixture of "- of the protein occurs at high surfactant to protein ratios. This
lactoglobulin and low methoxyl pectin (Ganzevles et al., 2006), mechanism is called the orogenic displacement (Mackie et al., 2000)
whey protein concentrate in the presence of sodium alginate or or replacement mechanism (Bos and Vliet, 2001). Courthaudon
&-carrageenan (Perez et al., 2009), "-lactoglobulin in the pres- et al. (1991) showed that the addition of a water-soluble non-
ence of sodium alginate or &-carrageenan (Perez et al., 2012), ionic surfactant to an oil-in-water emulsion stabilized by "-casein
hydroxypropylmethylcellulose (Camino et al., 2009), anionic or led to total displacement of the protein from the oil–water inter-
ionic surfactants and ovalbumin (Seta et al., 2012). face at the surfactant-to-protein molar ratio of 17 ± 1. Conversely,
The different adsorption mechanisms observed for low ionic surfactants tend to complex with the adsorbed protein. With
molecular weight surfactants and biopolymers also lead to dif- increasing surfactant concentrations, electrostatic interactions are
ferent molecular organization within the interfacial film. Small preponderant until all the protein charges are compensated by sur-
surfactants can pack together more closely at the interface than factant charges. The complex is then globally neutral and displays
biopolymers. The concentration of adsorbed surfactant molecules a higher surface activity. Upon further addition of surfactant, an
is typically 10−5 mol/m2 , which corresponds to adsorbed amounts increasing number of surfactant molecules interact with the com-
ranging between 1 and 2 mg/m2 (Bos and Vliet, 2001). For proteins, plex through hydrophobic interactions, making step-by-step the
the interfacial concentration is much lower: about 10−7 mol/m2 , complex more hydrophilic, and less surface active. Due to the com-
with an adsorbed amount ranging between 2 and 3 mg/m2 (Bos and petition in the adsorbed layer, the soluble complex is progressively
Vliet, 2001). For polysaccharides, owing to their higher molecular replaced by surfactant molecules, so that typically when the free
weight, the concentration of adsorbed molecules is even lower surfactant concentration in the bulk reaches the critical micellar
(about 10−8 mol/m2 ) whereas the adsorbed amount is higher concentration the adsorption layer is mainly formed by the small-
(2–14 mg/m2 ) (Buffo et al., 2001; Padala et al., 2009; Siew and surfactant molecules (solubilisation mechanism) (Bos and Vliet,
Williams, 2008). 2001; Maldonado-Vaderrama and Rodriguez Patino, 2010; Miller
372 E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378

et al., 2000). Compared to non-ionic surfactants, a much higher (a hydrophobic polyether–polysiloxane block copolymer) was used
concentration of ionic surfactant is required for complete displace- as the external emulsifier.
ment of a protein from the interface. This is attributable to the
relatively low surface activity of the charged surfactant and to 4.1.3. Comparison of emulsion stabilizing properties
the strong interfacial surfactant–protein interactions (Dickinson, Low molecular weight surfactants are more effective than
2001). Few studies are devoted to the interfacial behavior of mixed biopolymers at producing emulsions with small droplets – owing to
polysaccharide–surfactant systems and these systems also display kinetic and structural aspects, as previously mentioned. Obtaining
a complex behavior (Garti et al., 1999; Mezdour et al., 2008; Ropers globules with a small diameter and with a narrow size distribu-
et al., 2008). tion is an efficient way of reducing destabilization mechanisms
such as Ostwald ripening and creaming. However, the long-term
4.1.2. Comparison of emulsifying properties stability of an emulsion is mostly influenced by flocculation and
During the emulsification process, the adsorption kinetics of an coalescence. The biopolymer segments that remain suspended
amphiphilic molecule is influenced by its molecular weight: a small into the aqueous phase provide an efficient steric stabilization of
monomeric surfactant adsorbs at the interface very quickly, which the oil droplet. Indeed, the strength and the range of the repul-
contributes to the production of small droplets. Small-molecule sive interactions between two adjacent droplets are important
surfactants such as Tween 20 and sodium dodecyl sulphate are factors. The minimum interdroplet separation is primarily deter-
more effective at producing small droplets in nanoemulsions than mined by the physical space occupied by the adsorbed molecules,
caseinate and "-lactoglobulin under similar homogenization con- except at very low ionic strength. The relative thickness of the
ditions because they adsorb to the droplet surfaces more rapidly thin films between closely approaching oil droplets is classified
(Qian and McClements, 2011). Moreover, the small surfactant size in ascending order: monomeric surfactant molecules, proteins and
allows a higher droplet curvature (reduced steric hindrance) and hydrocolloids (Dickinson, 2009). Stability against droplet coales-
globules are generally smaller than with biopolymers. Biopolymer- cence also depends on the mechanical properties of the interfacial
stabilized emulsions generally have droplet sizes in the order film. In this respect again, biopolymers are more efficient than
of 1 !m or more and formation of nanoemulsions with them low molecular weight surfactants. For example, some high mole-
is thus more difficult. However very recently, McClement and cular weight biopolymers that slowly adsorb at the interface and/or
co-workers have obtained nanoemulsions with "-lactoglobulin decrease the interfacial tension to a small extent only, still confer
(Li et al., 2012; Qian and McClements, 2011; Qian et al., 2012), the interface very good viscoelastic properties allowing stabiliz-
sodium caseinate (Qian and McClements, 2011), and whey pro- ing emulsions (Bouyer et al., 2011). Moreover, the layer-by-layer
tein isolate (Lee et al., 2011). The emulsification process used often deposition method is an efficient way to protect globules against
limits the choice of emulsifiers that can be used in nanoemul- environmental stresses (heating, ionic strength, pH etc.). However,
sion formulation. For example formation of nanoemulsions by the stability of such emulsions is highly dependent on the amounts
phase inversion temperature or spontaneous emulsification meth- of the biopolymers used. Indeed, an incorrect ratio between the
ods is not possible with proteins or polysaccharides (McClements, layer components can lead to bridging flocculation (if the amount of
2011). As for microemulsions, to our knowledge, no successful biopolymer does not allow a full coverage of the globules) or deple-
formulation study has been reported with biopolymers, without tion flocculation (if there is an excess of unabsorbed biopolymer
addition of small-molecule surfactants. Actually, only low mole- or of unabsorbed soluble complexes). When the formulation and
cular weight surfactants allow reaching the ultra low interfacial process conditions are carefully controlled, it is possible to obtain
tensions at specific monolayer curvature required for the obtention emulsions with improved stability compared to those stabilized
of a microemulsion (McClements, 2012). When comparing biopoly- with small surfactants. From an application standpoint, another
mers to small surfactants, it is also important to note that due to interest of biopolymers is that thanks to their viscosifying pro-
their solubility in water, biopolymers are only used to produce perties, they simultaneously stabilize emulsions and control their
O/W emulsions when used alone. For instance, water-in-silicon oil texture. It is therefore possible to obtain emulsion gels i.e. soft-solid
emulsions cannot be stabilized by human serum albumin by itself materials which are both an emulsion and a gel (Benna-Zayani et al.,
and it is necessary to add an oil-soluble and end-functionalized 2008; Dickinson, 2012; Weiss et al., 2005).
silicon polymer (Bartzoka et al., 2000; Zelisko et al., 2008). Con-
versely, small surfactants, depending on their HLB (hydrophilic 4.2. Promising applications of biopolymers for the stabilization of
lipophilic balance), can be applied to the formulation of O/W or W/O pharmaceutical emulsions
emulsions. This also explains why double emulsions (W1 /O/W2 )
formulated with biopolymers contain an oil-soluble surfactant to Various biopolymer stabilization mechanisms have been
prepare the primary W1 /O emulsion. The oil-soluble surfactant described, demonstrating the possibility to formulate stable “clean
commonly used in food emulsions is PGPR (polyglycerol ester of labelled” emulsions. Indeed, it is possible to completely replace
polyricinoleic acid). The biopolymer can be incorporated in the pri- small-molecule surfactants by biopolymers to formulate O/W
mary emulsion, either to gelify the aqueous internal phase W1 as emulsions with increased stability. In double emulsions, the
with gelatine (Sapei et al., 2012) or to decrease the concentra- amount of small-molecule surfactants can only be reduced by the
tion of the oil-soluble surfactants as with sodium caseinate (Su use of biopolymers, as previously explained. In the modern context
et al., 2006). Biopolymers are also used as external emulsifiers: of health and environmental concerns and sustainable develop-
sodium caseinate (Su et al., 2006), sodium caseinate–dextran con- ment, biopolymers such as proteins and polysaccharides could thus
jugates (Fechner et al., 2007), a modified gum arabic (Su et al., be a good alternative to synthetic surfactants in some pharmaceu-
2008), xanthan + whey protein isolate (Benichou et al., 2007a) tical formulations, especially for oral and topical applications.
carboxymethylcellulose + whey protein isolate (Murillo-Martinez As previously described, various naturally occurring polymers
et al., 2011), low-methyl pectin + whey protein isolate (Murillo- can be used either alone or in combination for emulsion stabiliza-
Martinez et al., 2011), mesquite gum + maltodextrin + whey protein tion, and many of them are already used in the pharmaceutical
concentrate (Carillo-Navas et al., 2012), etc. O1 /W/O2 double emul- field for other applications. A wide variety of structures and tex-
sions have been less studied although Benichou et al. (2007b) have tures can be obtained with biopolymers: O/W macroemulsions,
reported that it was possible to stabilize the inner oil droplets O/W nanoemulsions, O/W/O or W/O/W double emulsions, and
with xanthan–whey–protein isolate hybrids, whereas Abil EM90 emulsion gels. Nanoemulsions, double emulsions, and emulsion
E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378 373

gels are particularly interesting systems from a pharmaceutical (2010) have shown that the physical stability of oil droplets coated
standpoint. Indeed, nanoemulsions exhibit some potential advan- by multilayer biopolymers is largely determined by the electrical
tages over conventional emulsions. They display a higher stability characteristics of the outer biopolymer layer. If a delivery system
especially to gravitational separation. They can also increase the must remain stable under acid conditions (i.e. in the stomach),
bioavailability of encapsulated lipophilic pharmaceutically active but breakdown at neutral pH (i.e. in the intestine), then a cationic
ingredients (Gao et al., 2011; Prakash and Thiagarajan, 2011). The outer coating of chitosan could be chosen. On another hand, if a
obtention of nanoemulsions formulated with biopolymers is thus delivery system must remain stable under neutral conditions but
a major breakthrough. McClements and co-workers (Qian et al., breakdown at acid pH, then an anionic outer layer of alginate or
2012) have shown that it is possible to obtain nanoemulsions pectin could be used. This shows that it is possible to achieve a
encapsulating "-carotene, a component that may reduce the inci- site-specific delivery of an oil-soluble compound contained in an
dence of some diseases (some cancers, cardiovascular diseases, emulsion. More recently, Benjamin et al. (2012) have shown that
age-related macular degeneration and cataracts). Interestingly, "- alternative layers of "-lactoblobulin and pectin around oil droplets
carotene degradation was considerably slower in "-lactoglobulin were very efficient to control the release of encapsulated volatile
stabilized nanoemulsions than in Tween-20 stabilized nanoemul- compounds using pH and salt as release triggers. A review on the
sions although further studies are still required to clearly identify use of structured emulsion-based delivery systems to control the
the mechanisms involved. It would thus be very interesting to study digestion and release of lipophilic components has recently been
the encapsulation and the release of pharmaceutically active ingre- written by McClements and Li (2010). This knowledge can be used
dients from such systems. As regards double emulsions, it was to select the most appropriate emulsion-based delivery system for
shown that complexes of whey protein isolate and xanthan gum specific pharmaceutical applications, such as encapsulation, con-
could stabilize W/O/W emulsions (as external emulsifiers) and act trolled digestion, and targeted release.
as an efficient barrier against the release of vitamin B1 entrapped in Grigoriev et al. (2008) also used the LBL technique to encap-
the inner aqueous droplets (Benichou et al., 2007a). In their study of sulate an emulsion. They performed a layer-by-layer deposition
an O/W/O double emulsion stabilized with whey protein isolate and on a liquid core using synthetic polymers such as poly(sodium
xanthan gum, Garti and co-workers (Benichou et al., 2007b) have 4-styrenesulfonate and poly(diallyldimethylammonium chloride)
shown that the use of these biopolymers to stabilize the primary to form loaded micro- and nano-containers. The droplet emul-
emulsion allows for a high yield of encapsulation of flumethrin® (a sion sizes were in the order of 2 !m. This novel approach is rather
veterinary drug model). Moreover, the absence of monomeric sur- general and could be extended to biopolymers and to the encapsu-
factant in the intermediate aqueous phase of the double emulsion lation of a broad range of pharmaceutical emulsions comprised of
prevents uncontrolled release of the entrapped compound through various hydrophobic substances and oil-soluble compounds.
the micellar diffusion mechanism. Recently, Murillo-Martinez et al.
(2011) have obtained edible films from W/O/W double emulsions 4.3. Challenging issues raised by biopolymers in the field of
stabilized by protein–polysaccharide complexes. The mechanical pharmaceutical emulsions
properties of these edible films were comparable to those obtained
from hydrophilic edible films of proteins or polysaccharides, but It is important to note that the use of natural polymers also raises
with vapour permeability comparable to that of lipid-made edi- some challenging issues that will have to be addressed before their
ble films. Those systems can encapsulate an active molecule, and current use in the pharmaceutical field.
constitute an attractive form that could prevent destabilization First, the variability of biopolymer composition depending on
upon long-term storage and microbacterial development (see next its source and origin is of importance and may be a problem for
section). Regarding emulsion gels, it should be possible to design the reproducibility of the formulations. Indeed, composition vari-
delivery systems with long-term physical stability (entrapment ations may influence the interfacial properties of biopolymers and
of the globules preventing coalescence, flocculation and cream- their rheological behaviors, among others. However, this issue has
ing) and with mucoadhesive properties. Indeed, the mucoadhesive already been addressed for some polymers such as HPMC, pectin,
properties of some biopolymers such as chitosan and hyaluronan gelatin, etc., which are already available in pharmaceutical grades.
are of particular interest for oral and topical routes of administra- Second, most of these polymers are currently used in the food
tion, including vaginal and rectal ones. industry and their ingestion is safe. However, the use of some nat-
The association of proteins and polysaccharides properties to ural polymers can raise allergenicity concerns. For instance, whey
stabilize emulsions is increasingly studied in order to enhance proteins such as "-lactoglobulin are well known to cause food aller-
emulsion stability. Depending on the association method of gies. The route of administration is important. Indeed, for topical
biopolymers (covalent, electrostatic, no interaction), the obtained use "-lactoglobulin might not induce allergic reactions, as it would
properties are different and open various promising application remain in the stratum corneum. Complementary studies are thus
perspectives. Covalent and electrostatic protein–polysaccharide required. Moreover, Mouécoucou et al. (2004) have shown that the
associations protect emulsions against destabilization under unfa- protein to polysaccharide association is likely to reduce the IgG/IgE
vorable environmental conditions (i.e. heating, freezing, high binding of the protein and thus its allergenicity. In addition, in the
electrolyte concentrations, lipid oxidation, etc.) (Aoki et al., 2005; framework of topical administration, the high molecular weight of
Gu et al., 2005, 2007; Guzey and McClements, 2006a; O’Regan the protein–polysaccharide complex could also be an advantage to
and Mulvihill, 2010). The most promising stabilization technique prevent its skin permeation. Other polymers, such as chitosan or
is most probably the layer-by-layer deposition. This method offers hyaluronan, will need to be further studied and evaluated as to oral
many perspectives regarding the targeted delivery of oil soluble ingestion safety. The safety of chitosan for its use in food has already
compounds. A potential benefit of multilayer emulsions as deli- been recognized in the U.S. (Baldrick, 2010). These safety and risk-
very systems is that they can be formulated entirely from natural benefit preoccupations will also have to be taken into account for
excipients (oil, proteins, polysaccharides) using simple processing other delivery routes (dermatologic, ocular, colonic, rectal, etc.)
operations (homogenization, mixing). These biopolymer coatings when these natural biopolymers are part of formulations.
can be designed to improve the stability of delivery systems to Third, as already mentioned, the obtained droplet sizes seem
environmental stresses, to protect encapsulated compound from generally to be in the order of 1 !m or more with either broad
chemical degradation, and to release encapsulated components in or narrow distributions. This will necessarily also influence the
response to specific environmental triggers. For example, Li et al. delivery route, as it would not be safe to use them for parenteral
374 E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378

administration (except for the newly developed nanoemulsions). membranes produced by layer-by-layer electrostatic deposition technique.
For this reason, biopolymer stabilized emulsions seem to be better Food Hydrocolloids 19, 209–220.
Avadi, M.R., Sadeghi, A.M.M., Mohammadpour, N., Abedin, S., Atyabi, F., Dinar-
designed for oral or topical administration. vand, R., Rafiee-Tehrani, M., 2010. Preparation and characterization of insulin
Finally, because of their organic origin, there are bacteriological nanoparticles using chitosan and Arabic gum with ionic gelation method.
and storage concerns. The addition of antimicrobial agents may be Nanomedicine: Nanotechnol. Biol. Med. 6, 58–63.
Baldrick, P., 2010. The safety of chitosan as a pharmaceutical excipient. Regul. Toxi-
necessary. For example, benzoic acid, sodium benzoate, methyl- col. Pharmacol. 56, 290–299.
and propyl-parabens are currently used in the pharmaceutical field Banker, G., Peck, G., Jan, S., Pirakitikulr, P., 1981. Evaluation of hydroxyl-propyl cel-
(Kibbe, 2006a). However, the use of these additives is incompa- lulose and hydroxypropyl methylcellulose as aqueous based film coatings. Drug
Dev. Ind. Pharm. 7, 693–716.
tible with respect to sustainable development concerns. Another
Bantchev, G.B., Schwartz, D.K., 2003. Surface shear rheology of "-casein layers at the
solution consists in lyophilizing (Corveleyn and Remon, 1998; Li air/solution interface: formation of a two-dimensional physical gel. Langmuir
et al., 2008) or spray-drying (Gharsallaoui et al., 2007, 2010a, 2012) 19, 2673–2682.
Bartzoka, V., Chan, G., Brook, M.A., 2000. Protein–silicone synergism at liquid/liquid
emulsions to avoid their destabilization upon long-term storage.
interfaces. Langmuir 16, 4589–4593.
Starches, gum arabic, whey-protein isolate, gelatin, pectins, and Becher, P., 1983. Encyclopedia of Emulsion Technology, vol. 2. Marcel Dekker, New
their associations have shown suitable properties for multi-layered York.
shell formation by spray-drying (Gharsallaoui et al., 2007). This Benichou, A., Aserin, A., Garti, N., 2007a. W/O/W double emulsions stabilized with
WPI-polysaccharide complexes. Colloids Surf. A: Physicochem. Eng. Aspects 294,
process can thus enable the encapsulation of oil-soluble phar- 20–32.
maceutical compounds. However, heating and evaporation of the Benichou, A., Aserin, A., Garti, N., 2007b. O/W/O double emulsions stabilized with
solvent necessary for the sedimentation/crystallization of poly- WPI–polysaccharide conjugates. Colloids Surf. A: Physicochem. Eng. Aspects
297, 211–220.
mers layers may also cause chemical transformations in the core. Benjamin, O., Silcock, P., Leus, M., Everett, D.W., 2012. Multilayer emulsions as deliv-
That is why another currently studied method is the formation of ery systems for controlled release of volatile compounds using pH and salt
the solid layers around the oil droplets by chemical reactions. Low triggers. Food Hydrocolloids 27, 109–118.
Benmouffok-Benbelkacem, G., Caton, F., Baravian, C., Skali-Lami, S., 2010. Non-linear
methoxy pectin microcapsules with hydrophobic core can thus be viscoelasticity and temporal behaviour of typical yield stress fluids: Carbopol,
formed by interfacial reaction (Muhiddinov et al., 2004). In this Xanthan and Ketchup. Rheol. Acta 49, 305–314.
study, the authors promoted interfacial film formation by using Benna-Zayani, M., Kbir-Ariguib, N., Trabelsi-Ayadi, M., Grossiord, J.L., 2008. Stabi-
lization of W/O/W double emulsion by polysaccharides as weak gels. Colloids
either anionic sodium dodecyl sulphate (physical cross-linking
Surf. A: Physicochem. Eng. Aspects 316, 46–54.
interaction) or cationic benzalkonium chloride (electrostatic Billa, N., Yuen, K.H., Khader, M.A., Omar, A., 2000. Gamma scintigraphic study
interaction). The obtained microcapsules were smaller than 3 !m. of the gastrointestinal transit and in vivo dissolution of a controlled release
diclofenac sodium formulation in xanthan gum matrices. Int. J. Pharma. 201,
Again, this approach could be extended to biopolymers.
109–120.
Binks, B.P., Clint, J.H., Dyab, A.K.F., Fletcher, P.D.I., Kirkland, M., Whitby, C., 2003.
5. Conclusion Ellipsometric study of monodisperse silica particles at an oil–water interface.
Langmuir 19, 8888–8893.
The growing interest and demand of consumers in natural pro- Binks, B.P., Fletcher, P.D.I., Paunov, V.N., Segal, D., 2000. Equilibrium and dynamic
adsorption of C12 E5 at the air–water surface investigated using ellipsometry and
ducts incites industrials to invest in this direction. The present tensiometry. Langmuir 16, 8926–8931.
review reveals the potentialities of biopolymers for the formula- Blijdenstein, T.B.J., van der Linden, E., van Vliet, T., van Aken, G.A., 2004. Scaling
tion and stabilization of pharmaceutical emulsions. Indeed, due to behavior of delayed demixing, rheology, and microstructure of emulsions floc-
culated by depletion and bridging. Langmuir 20, 11321–11328.
the versatility and large number of available biopolymers, protein Bonnet, C., Corredig, M., Alexander, M., 2005. Stabilization of caseinate covered oil
and/or polysaccharide stabilized emulsions appear to be a very droplets during acidification with high methoxyl pectin. J. Agric. Food Chem. 53,
promising approach for pharmaceutics in the same way as they 8600–8606.
Bos, M.A., Vliet, T.V., 2001. Interfacial rheological properties of adsorbed protein
already are in the food industry. However, further research work layers and surfactants: a review. Adv. Colloid Interface Sci. 91, 437–471.
is still needed to meet the specific requirements of this new field Bouyer, E., Mekhloufi, G., Le Potier, I., du Fou de Kerdaniel, T., Grossiord, J.-L., Rosilio,
of applications. For examples, formulation feasibility, interactions V., Agnely, F., 2011. Stabilization mechanism of oil-in-water emulsions by "-
lactoglobulin and gum arabic. J. Colloid Interface Sci. 354, 467–477.
with active molecules, and possible influences of biopolymers on
Bradford, M., 1976. A rapid and sensitive method for the quantification of micro-
drug bioavailability and pharmacokinetics must be studied. It is gram quantities of protein utilizing the principle of protein-dye binding. Anal.
worth to note that the formulation and stability control of the Biochem. 72, 248–254.
obtained emulsions require a thorough knowledge of the phys- Brime, B., Moreno, M.A., Frutos, G., Ballesteros, M.P., Frutos, P., 2002. Amphotericin
B in oil–water lecithin-based microemulsions: formulation and toxicity evalu-
ical chemistry of these systems. Recent techniques have led to ation. J. Pharma. Sci. 91, 1178–1185.
great progress in understanding the mechanisms of stabilization Brittain, H.G., 2001. Particle-size distribution, Part I: representation of particle shape
and consequently in optimizing the formulation process. and size. Pharma. Technol. 25, 38–45.
Brummer, Y., Cui, W., Wang, Q., 2003. Extraction, purification and physicochemical
characterization of fenugreek gum. Food Hydrocolloids 17, 229–236.
Acknowledgement Buffo, R.A., Reineccius, G.A., Oehlert, G.W., 2001. Factors affecting the emulsifying
and rheological proprerties of gum acacia in beverage emulsions. Food Hydro-
We thank Dr. Christine Vauthier for her useful remarks and colloids 15, 53–66.
Buggins, T.R., Dickinson, P.A., Taylor, G., 2007. The effects of pharmaceutical excipi-
advices for the review redaction. ents on drug disposition. Adv. Drug Deliv. Rev. 59, 1482–1503.
Burgess, D.J., Ozlen Sahin, N., 1997. Interfacial rheological and tension properties of
References protein films. J. Colloid Interface Sci. 189, 74–82.
Bylaite, E., Nylander, T., Venskutonis, R., Jönsson, B., 2001. Emulsification of caraway
Agboola, S.O., Dalgleish, D.G., 1995. Calcium induced destabilization of oil-in-water essential oil in water by lecithin and "-lactoglobulin: emulsion stability and
emulsions stabilized by caseinate or by "-lactoglobulin. J. Food Sci. 60, 399–404. properties of the formed oil-aqueous interface. Colloids Surf. B: Biointerf. 20,
Akhtar, M., Dickinson, E., 2007. Whey protein–maltodextrin conjugates as emulsi- 327–340.
fying agents: an alternative to gum arabic. Food Hydrocolloids 21, 607–616. Camino, N.A., Pérez, O.E., Carrera Sanchez, C., Rodriguez Patino, J.M., Pilosof,
Alderman, D.A., 1984. A review of cellulose ethers in hydrophilic matrices for oral A.M.R., 2009. Hydroxypropylmethylcellulose surface activity at equilibrium and
controlled-release dosage forms. Int. J. Pharma. Technol. Prod. Manuf. 5, 1–9. adsorption dynamics at the air–water and the oil–water interfaces. Food Hydro-
Ambrosio, L., Borzacchiello, A., Netti, P.A., Nicolais, L., 1999. Rheological study colloids 23, 2359–2368.
on hyaluronic acid and its derivative solutions. J. Macromol. Sci., Part A 36, Capek, I., 2004. Degradation of kinetically-stable o/w emulsions. Adv. Colloid Inter-
991–1000. face Sci. 107, 125–155.
Anderson, D.M.W., Bridgeman, M.M.E., Farquhar, J.G.K., McNab, C.G.A., 1983. The Carillo-Navas, H., Cruz-Olivares, J., Varela-Guerrero, V., Alamilla-Beltran, L., Vernon-
chemical characterization of the test article used in toxicological studies of gum Carter, E.J., Pérez-Alonso, C., 2012. Rheological properties of a double emulsion
arabic (Acacia senegal (L) Willd). Int. Tree Crops J. 2, 245–254. nutraceutical system incorporating chia essentials oil and ascorbic acid sta-
Aoki, T., Decker, E.A., McClements, D.J., 2005. Influence of environmental stresses bilized by carbohydrate polymer–protein blends. Carbohydr. Polymers 87,
on stability of O/W emulsions containing droplets stabilized by multilayered 1231–1235.
E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378 375

Casas, J.A., Santos, V.E., Garcia-Ochoa, F., 2000. Xanthan gum production under Dickinson, E., 2009. Hydrocolloids as emulsifiers and emulsion stabilizers. Food
several operational conditions: molecular structure and rheological properties. Hydrocolloids 23, 1473–1482.
Enzyme Microb. Technol. 26, 282–291. Dickinson, E., 2011. Mixed biopolymers at interfaces: competitive adsorption and
Castellani, O., Gaillard, C., Vié, V., Al-Assaf, S., Axelos, M., Phillips, G.O., Anton, M., multilayer structures. Food Hydrocolloids 25, 1966–1983.
2010. Hydrocolloids with emulsifying capacity. Part 3 – adsorption and struc- Dickinson, E., 2012. Emulsion gels: the structuring of soft solids with protein-
tural properties at the air-water surface. Food Hydrocolloids 24, 131–141. stabilized oil droplets. Food Hydrocolloids 28, 224–241.
Ceulemans, J., Vinckier, I., Ludwig, A., 2002. The use of xanthan gum in an ophthalmic Dickinson, E., Davies, E., 1999. Influence of ionic calcium on stability of sodium
liquid dosage form: rheological characterization of the interaction with mucin. caseinate emulsions. Colloids Surf. B: Biointerf. 12, 203–212.
J. Pharma. Sci. 91, 1117–1127. Dickinson, E., Euston, S.R., Woskett, C.M., 1990. Competitive adsorption of food
Chanamai, R., McClements, D.J., 2002. Comparison of gum arabic, modified starch, macromolecules and surfactants at the oil-water interface. Prog. Colloid Polymer
and whey protein isolate as emulsifiers: influence of pH, CaCl2 and temperature. Sci. 82, 65–75.
J. Food Sci. 67, 120–125. Dickinson, E., Galaska, V.B., 1991. Emulsion stabilization by ionic and covalent com-
Chatelier, R.C., Minton, A.P., 1996. Adsorption of globular proteins on locally pla- plexes of "-lactoglobulin with polysaccharides. Food Hydrocolloids 5, 281–296.
nar surfaces: models for the effect of excluded surface area and aggregation of Dickinson, E., Matsumara, Y., 1994. Proteins at liquid interfaces: role of the molten
adsorbed protein on adsorption equilibria. Biophys. J. 71, 2367–2374. globule state. Colloids Surf. B: Biointerf. 3, 1–17.
Chhabra, R.P., 1993. Bubbles Drops and Particles in Non-Newtonian Fluids. CRC Press, Dickinson, E., Pawlowsky, K., 1998. Influence of $-carrageenan on the properties of
Boca Raton, FL. a protein-stabilized emulsion. Food Hydrocolloids 12, 417–423.
Chowhan, Z.T., 1980. Role of binders in moisture-induced hardness increase in com- Doyle, J.P., Giannouli, P., Rudolph, B., Morris, E.R., 2010. Preparation, authentica-
pressed tablets and its effect on in vitro disintegration and dissolution. J. Pharma. tion, rheology and conformation of theta carrageenan. Carbohydr. Polymers 80,
Sci. 69, 1–4. 648–654.
Chuah, A.M., Kuroiwa, T., Kobayashi, I., Nakajima, M., 2009. Effect of chitosan on the Ducel, V., Richard, J., Popineau, Y., Boury, F., 2004a. Adsorption kinetics and rhe-
stability and properties of modified lecithin stabilized oil-in-water monodis- ological interfacial properties of plant proteins at the oil–water interface.
perse emulsion prepared by microchannel emulsification. Food Hydrocolloids Biomacromolecules 5, 2088–2093.
23, 600–610. Ducel, V., Richard, J., Saulnier, P., Popineau, Y., Boury, F., 2004b. Evidence and charac-
Ciancia, M., Noseda, M.D., Matulewicz, M.C., Cerezo, A.S., 1993. Alkali-modification of terization of complex coacervates containing plant proteins: application to the
carrageenans: mechanism and kinetics in the kappa/iota-, mu/nu- and lambda- microencapsulation of oil droplets. Colloids Surf. A: Physicochem. Eng. Aspects
series. Carbohydr. Polymers 20, 95–98. 232, 239–247.
Claesson, P.M., Ninham, B.W., 1992. pH-dependent interaction between adsorbed Dulong, V., Lack, S., Le Cerf, D., Picton, L., Vannier, J.P., Muller, G., 2004.
chitosan layers. Langmuir 8, 1406–1412. Hyaluronan-based hydrogels particles prepared by crosslinking with trisodium
Cook, W., 2006. Pectin. In: Rowe, R.C., Sheskey, P.J., Owen, S.C. (Eds.), Handbook of trimetaphosphate. Synthesis and characterization. Carbohydr. Polymers 57, 1–6.
Pharmaceutical Excipients. , 5th ed. Pharmaceutical Press and American Phar- Dunstan, D.E., Chen, Y., Liao, M.-L., Salvatore, R., Boger, D.V., Prica, M., 2001. Structure
macist Association, London, pp. 507–508. and rheology of the $-carrageenan/locust bean gum gels. Food Hydrocolloids 15,
Copelin, C., 1998. Practical points in the use of albumin for hypovolemia. J. Peri 475–484.
Anesth. Nurs. 13, 118–120. Effendy, I., Maibach, H.I., 1995. Surfactants and experimental irritant contact der-
Corveleyn, S., Remon, J.P., 1998. Formulation of a lyophilized dry emulsion tablet for matitis. Contact Dermat. 33, 217–225.
the delivery of poorly soluble drugs. Int. J. Pharma. 166, 65–74. El-Gazayerly, O.N., 2003. Release of pentoxifylline from xanthan gum matrix tablets.
Courthaudon, J.-L., Dickinson, E., Dalgleish, D.G., 1991. Competitive adsorption of Drug Dev. Ind. Pharma. 29, 241–246.
"-casein and nonionic surfactants in oil-in-water emulsions. J. Colloid Interface Erni, P., Windhab, E.J., Gunde, R., Graber, M., Pfister, B., Parker, A., Fischer, P., 2007.
Sci. 145, 390–395. Interfacial rheology of surface-active biopolymers: Acacia senegal gum versus
Coviello, T., Alhaique, F., Dorigo, A., Matricardi, P., Grassi, M., 2007. Two galactoman- hydrophobically modified starch. Biomacromolecules 8, 3458–3466.
nans and scleroglucan as matrices for drug delivery: preparation and release Euston, S.R., Finnigan, S.R., Hirst, R.L., 2000. Aggregation kinetics of heated whey
studies. Eur. J. Pharma. Biopharma. 66, 200–209. protein-stabilized emulsions. Food Hydrocolloids 14, 155–161.
Cross, M.M., 1965. Rheology of non Newtonian fluids: a new flow equation for Fabra, M.J., Talens, P., Chiralt, A., 2008. Effect of alginate and &-carrageenan on tensile
pseudoelastic systems. J. Colloid Sci. 20, 417–437. properties and water vapor permeability of sodium caseinate-lipid based films.
Cserháti, T., Forgács, E., Oros, G., 2002. Biological activity and environmental impact Carbohydr. Polymers 74, 419–426.
of anionic surfactants. Environ. Int. 28, 337–348. Fauconnier, M.L., Blecker, C., Groyne, J., Razafindralambo, H., Vanzeveren, E., Marlier,
Dahl, T.C., Calderwood, T., Bormeth, A., Trimble, K., 1990. Influence of physicochem- M., Paquot, M., 2000. Characterization of two Acacia gums and their fractions
ical properties of hydroxypropyl methylcellulose on naproxen release from using a Langmuir film balance. J. Agric. Food Chem. 48, 2709–2712.
sustained release matrix tablets. J. Control. Release 14, 1–10. Fechner, A., Knoth, A., Scherze, I., Muschiolik, G., 2007. Stability and release prop-
Dalgleish, D.G., 1997. Adsorption of protein and the stability of emulsions. Trends erties of double-emulsions stabilised by caseinate-dextran conjugates. Food
Food Sci. Technol. 8, 1–6. Hydrocolloids 21, 943–952.
Damodaran, S., 2005. Protein stabilization of emulsions and foams. J. Food Sci. 70, Ferri, J.K., Kotsmar, C., Miller, R., 2010. From surfactant adsorption kinetics to asym-
54–66. metric nanomembrane mechanics: pendent drop experiments with subphase
Daniels, R., Barta, A., 1993. Preparation, characterization and stability assessment exchange. Adv. Colloid Interface Sci. 161, 29–47.
of oil-in-water emulsions with hydroxypropylmethylcellulose as emulsifier. In: Friberg, S.E., Venable, R.L., 1983. Microemulsions. In: Becher, P. (Ed.), Encyclopedia
Proceedings of Pharmaceutics Technology Conference, Elsinore, pp. 51–60. of Emulsion Technology, vol. 1. Marcel Dekker, New York, pp. 287–336.
Day, A.J., Prestwich, G.D., 2002. Hyaluronan-binding proteins: tying up the giant. J. Gacesa, P., 1988. Alginates. Carbohydr. Polymers 8, 161–182.
Biol. Chem. 277, 4585–4588. Ganzevles, R.A., Cohen Stuart, M.A., van Vliet, T., de Jongh, H.H.J., 2006. Use of
Day, A.J., Sheehan, J.K., 2001. Hyaluronan: polysaccharide chaos to protein organi- polysaccharides to control protein adsorption at the air–water interface. Food
zation. Curr. Opin. Struct. Biol. 11, 617–622. Hydrocolloids 20, 872–878.
Dea, I.C.M., Morrison, A., 1975. Chemistry and interactions of seed galactomannans. Gao, F., Zhang, Z., Bu, H., Huang, Y., Gao, Z., Shen, J., Zhao, C., Li, Y., 2011. Nanoemulsion
Adv. Carbohydr. Chem. Biochem. 31, 241–312. improves the oral absorption of candersartan cilexetil in rats: performance and
Demé, B., Rosilio, V., Baszkin, A., 1995. Polysaccharides at interfaces. 1. Adsorption of mechanism. J. Control. Release 149, 168–174.
cholesteryl-pullulan derivatives at the solution–air interface. Kinetic study by Garg, H.G., Hales, C.A., 2004. Chemistry and Biology of Hyaluronan. Elsevier Science
surface tension measurements. Colloids Surf. B: Biointerf. 4, 357–365. and Technology Books, London.
Demitriades, K., Coupland, J.N., McClements, D.J., 1997. Physico-chemical properties Garti, N., 1999a. Hydrocolloids as emulsifying agents for oil-in-water emulsions. J.
of whey protein-stabilized emulsions as affected by heating and ionic strength. Dispers. Sci. Technol. 20, 327–355.
J. Food Sci. 62, 462–467. Garti, N., 1999b. What can nature offer from an emulsifier point of view: trends and
Dickinson, E., 1995. Emulsion stabilization by polysaccharides and progress? Colloids Surf. A: Physicochem. Eng. Aspects 152, 125–146.
protein–polysaccharide complexes. In: Stephen, A.M. (Ed.), Food Polysac- Garti, N., Aserin, A., Slavin, Y., 1999. Competitive adsorption in O/W emulsions stabi-
charides and Their Applications. Marcel Dekker, New York, pp. 501–515. lized by the new Portulaca oleracea hydrocolloid and non ionic emulsifiers. Food
Dickinson, E., 1997. Properties of emulsions stabilized with milk proteins: overview Hydrocolloids 13, 139–144.
of some recent developments. J. Dairy Sci. 80, 2607–2619. Garti, N., Reichman, D., 1993. Hydrocolloids as food emulsifiers and stabilizers. Food
Dickinson, E., 1999. Adsorbed protein layers at fluid interfaces: interactions, struc- Microstruct. 12, 411–426.
ture and surface rheology. Colloids Surf. B: Biointerf. 15, 161–176. Garti, N., Reichman, D., 1994. Surface properties and emulsification activity of galac-
Dickinson, E., 2001. Milk protein interfacial layers and the relationship to emulsion tomannans. Food Hydrocolloids 8, 155–173.
stability and rheology. Colloids Surf. B: Biointerf. 20, 197–210. Gau, C.-S., Yu, H., Zografi, G., 1994. Surface viscoelasticity of "-casein monolayers at
Dickinson, E., 2003. Hydrocolloids at interfaces and the influence on the properties the air/water interface by Electrocapillary Wave diffraction. J. Colloid Interface
of dispersed systems. Food Hydrocolloids 17, 25–39. Sci. 162, 214–221.
Dickinson, E., 2006. Colloid science of mixed ingredients. Soft Matter 2, 642–652. Gharsallaoui, A., Roudaut, G., Chambin, O., Voilley, A., Saurel, R., 2007. Applications
Dickinson, E., 2008a. Emulsification and emulsion stabilization with of spray-drying in microencapsulation of food ingredients: an overview. Food
protein–polysaccharide complexes. In: Williams, P.A., Phillips, G.O. (Eds.), Res. Int. 40, 1107–1121.
Gums and stabilizers for the food industry-14. Royal Society, Cambridge UK, Gharsallaoui, A., Roudaut, G., Chambin, O., Voilley, A., Saurel, R., 2012. Pro-
pp. 221–232. perties of spray-dried food flavours microencapsulated with two-layered
Dickinson, E., 2008b. Interfacial structure and stability of food emulsions as affected membranes:roles of interfacial interactions and water. Food Chem. 132,
by protein–polysaccharide interactions. Soft Matter 4, 932–942. 1713–1720.
376 E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378

Gharsallaoui, A., Saurel, R., Chambin, O., Cases, E., Voilley, A., Cayot, P., 2010a. Utili- Jones, D.S., Mawhinney, H.J., 2006. Chitosan. In: Rowe, R.C., Sheskey, P.J., Owen, S.C.
sation of pectin coating to enhance spray-dry stability of pea protein-stabilized (Eds.), Handbook of Pharmaceutical Excipients. , 5th ed. Pharmaceutical Press
oil-in-water emulsions. Food Chem. 122, 447–454. and American Pharmacist Association, London, pp. 159–162.
Gharsallaoui, A., Yamauchi, K., Chambin, O., Cases, E., Saurel, R., 2010b. Effect of high Jones, R.T., 2004. Gelat manufacture and physio-chemical. In: Podczeck, F., Jones, B.E.
methoxyl pectin on pea protein in aqueous solution and at oil/water interface. (Eds.), Pharmaceutical Capsules. , 2nd edition. Pharmaceutical Press, London, pp.
Carbohydr. Polymers 80, 817–827. 23–60.
Glickman, M., 1982. Food Hydrocolloids. CRC Press, Boca Raton, FL. Jossic, L., Magnin, A., 2005. Structuring of gelled suspensions flowing through a sud-
Gomez-Diaz, D., Navata, J.M., 2004. Rheology of food stabilizers blend. J. Food Eng. den three dimensional expansion. J. Non-Newtonian Fluid Mech. 127, 201–212.
64, 143–149. Jourdain, J., Leser, M.E., Schmitt, C., Michel, M., Dickinson, E., 2008. Stability of
Graham, D.E., Phillips, M.C., 1979. Proteins at liquid interfaces. J. Colloid Interface emulsions containing sodium caseinate and dextran sulfate: relationship to
Sci. 70, 403–414 (415–427, 427–439). complexation in solution. Food Hydrocolloids 22, 647–659.
Grigoriev, D.O., Bukreeva, T., Möhwald, H., Shchukin, D.G., 2008. New method Kato, A., Sasaki, Y., Furuta, R., Kobayashi, K., 1990. Functional protein/polysaccharide
for fabrication of loaded micro- and nanocontainers: emulsion encapsulation conjugate prepared by controlled dry heating of ovalbumin/dextran mixtures.
by polyelectrolyte layer-by-layer deposition on the liquid core. Langmuir 24, Agric. Biol. Chem. 54, 107–112.
999–1004. Khalloufi, S., Alexander, M., Goff, H.D., Corredig, M., 2008. Physicochemical proper-
Grigoriev, D.O., Miller, R., 2009. Mono- and multilayer covered drops as carriers. ties of whey protein isolate stabilized oil-in-water emulsions when mixed with
Curr. Opin. Colloid Interface Sci. 14, 48–59. flaxseed gum at neutral pH. Food Res. Int. 41, 964–972.
Groeneweg, F., van Voorst Vader, F., Agterof, W.G.M., 1993. The dynamic stability of Khalloufi, S., Corredig, M., Goff, H.D., Alexander, M., 2009a. Flaxseed gums and their
w/o emulsions prepared with vegetable oil. Chemi. Eng. Sci. 48, 229–238. adsorption on whey protein-stabilized oil-in-water emulsions. Food Hydrocol-
Grossiord, J.L., Sellier, M., 2001. W/O/W multiple emulsions: a review of the release loids 23, 611–618.
mechanisms by break-up of the oily membrane. STP Pharma. Sci. 11, 331–339. Khalloufi, S., Corredig, M., Alexander, M., 2009b. Interactions between flaxseed gums
Grossiord, J.L., Stambouli, M., 2008. Potentialities of W/O/W multiple emulsions in and WPI-stabilized emulsion droplets assessed in situ using diffusing wave spec-
drug delivery and detoxification. In: Aserin, A. (Ed.), Multiple Emulsions Tech- troscopy. Colloids Surf. B: Biointerf. 68, 145–153.
nology and Applications. John Wiley and Sons, Inc, New Jersey, pp. 209–234. Kibbe, A.H., 2006a. Acacia. In: Rowe, R.C., Sheskey, P.J., Owen, S.C. (Eds.), Hand-
Gu, Y.S., Decker, E.A., McClements, D.J., 2005. Influence of pH and carrageenan type book of Pharmaceutical Excipients. , 5th ed. Pharmaceutical Press and American
on properties of "-lactoglobulin stabilized oil-in-water emulsions. Food Hydro- Pharmacist Association, London, pp. 1–3.
colloids 19, 83–91. Kibbe, A.H., 2006b. Guar Gum. In: Rowe, R.C., Sheskey, P.J., Owen, S.C. (Eds.), Hand-
Gu, Y.S., Decker, E.A., McClements, D.J., 2007. Application of multi-component book of Pharmaceutical Excipients. , 5th ed. Pharmaceutical Press and American
biopolymer layers to improve the freeze-thaw stability of oil-in-water emul- Pharmacist Association, London, pp. 315–317.
sions: "-lactoglobulin-%-carrageenan-gelatin. J. Food Eng. 80, 1246–1254. Kim, H.J., Decker, E.A., McClements, D.J., 2002. Role of postadsorption conforma-
Gueguen, J., Chevalier, M., Barbot, J., Schaeffer, F., 1988. Dissociation and aggregation tion changes of beta-lactoglobulin on its ability to stabilize oil droplets against
of pea legumin induced by pH and ionic strength. J. Sci. Food Agric. 44, 167–182. flocculation during heating at neutral pH. Langmuir 18, 7577–7583.
Guzey, D., Kim, H.J., McClements, D.J., 2004. Factors influencing the production of Klein, M., Aserin, A., Svitov, I., Garti, N., 2010. Enhanced stabilization of cloudy emul-
o/w emulsions stabilized by "-lactoglobulin-pectine membranes. Food Hydro- sions with gum Arabic and whey protein isolate. Colloids Surf. B: Biointerf. 77,
colloids 18, 967–975. 75–81.
Guzey, D., McClements, D.J., 2006a. Formation, stability and properties of multilayer Kogan, A., Garti, N., 2006. Microemulsions as transdermal drug delivery vehicles.
emulsions for application in the food industry. Adv. Colloid Interface Sci. 128, Adv. Colloid Interface Sci. 123-126, 369–385.
227–248. Kovacs, P., 1973. Useful incompatibility of xanthan gum with galactomannans. Food
Guzey, D., McClements, D.J., 2006b. Influence of environmental stresses on Technol. 27, 26–30.
O/W emulsions stabilized by beta-lactoglobulin-pectin and beta-lactoglobulin- Krägel, J., Derkatch, S.R., 2010. Interfacial shear rheology. Curr. Opin. Colloid Interface
pectin-chitosan membranes produced by the electrostatic layer-by-layer Sci. 15, 246–255.
deposition technique. Food Biophys. 1, 30–40. Kravtchenko, T.P.,1997. Application of acacia gum as a natural source of soluble
Hambleton, A., Debeaufort, F., Bonnotte, A., Voilley, A., 2009a. Influence of algi- dietary fibre. In: Food Ingredients Europe, Conference Proceedings. Miller Free-
nate emulsion-based films structure on its barrier properties and on the man, Maarssen (The Netherlands), pp. 56–60.
protection of microencapsulated aroma compound. Food Hydrocolloids 23, Kubo, W., Miyazaki, S., Dairaku, M., Togashi, M., Mikami, R., Attwood, D., 2004. Oral
2116–2124. sustained delivery of ambroxol from in situ gelling pectin formulations. Int. J.
Hambleton, A., Fabra, M.-J., Debeaufort, F., Dury-Brun, C., Voilley, A., 2009b. Interface Pharma. 271, 233–240.
and aroma barrier properties of iota-carrageenan emulsion-based films used for Kuo, J.W., 2006. Practical Aspects of Hyaluronan-Based Medical Products. CRC Press,
encapsulation of active food compounds. J. Food Eng. 93, 80–88. New York.
Harwood, R.J., 2006. Hypromellose. In: Rowe, R.C., Sheskey, P.J., Owen, S.C. (Eds.), Lajavardi, L., Camelo, S., Agnely, F., Luo, W., Goldenberg, B., Naud, M.-C., Behar-Cohen,
Handbook of Pharmaceutical Excipients. , 5th ed. Pharmaceutical Press and F., de Kozac, Y., Bochot, A., 2009. New formulation of vasoactive intestinal pep-
American Pharmacist Association, London, pp. 346–349. tide using liposomes in hyaluronic acid gel for uveitis. J. Control. Release 139,
Haug, A., 1964. Composition and properties of alginates. Thesis. Norwegian Institute 22–30.
of Technology, Trondheim. Laplante, S., Turgeon, S.L., Paquin, P., 2005. Effect of pH, ionic strength, and composi-
Hejazi, R., Amiji, M., 2003. Chitosan-based gastrointestinal delivery systems. J. Con- tion on emulsion stabilising properties of chitosan in a model system containing
trol. Release 89, 151–165. whey protein isolate. Food Hydrocolloids 19, 721–729.
Hemar, Y., Tamehana, M., Munro, P.A., Singh, H., 2001. Influence of xanthan gum Laplante, S., Turgeon, S.L., Paquin, P., 2006. Emulsion-stabilizing properties of chi-
on stability of sodium caseinate oil-in-water emulsions. Food Hydrocolloids 15, tosan in the presence of whey protein isolate: effect of the mixture ratio, ionic
513–519. strength and pH. Carbohydr. Polymers 65, 479–487.
Henni, W., Deyme, M., Stchakovsky, M., Le Cerf, D., Picton, L., Rosilio, V., 2005. Laza-Knoerr, A., Huang, N., Grossiord, J.-L., Couvreur, P., Gref, R., 2011. Interfacial
Aggregation of hydrophobically modified polysaccharides in solution and at the rheology as a tool to study the potential of cyclodextrin polymers to stabilize
air–water interface. J. Colloid Interface Sci. 281, 316–324. oil–water interfaces. J. Inclus. Phenom. Macrocycl. Chem. 69, 475–479.
Hennock, M., Rahalkar, R.R., Richmond, P., 1984. Effect of xanthan gum upon the Ledward, D.A., 1986. Gelation of gelatin. In: Mitchel, J.R., Ledward, D.A. (Eds.), Func-
rheology and stability of oil-water emulsions. J. Food Sci. 49, 1271–1274. tional Properties of Food Macromolecules. Elsevier, London.
Hino, T., Yamamoto, A., Shimabayashi, S., Tanaka, M., Tsujii, D., 2000. Drug release Lee, S.J., Choi, S.J., Li, Y., Decker, E.A., McClements, D.J., 2011. Protein-stabilized
from w/o/w emulsions prepared with different chitosan salts and concomitant nanoemulsions and emulsions: comparison of physicochemical stability, lipid
creaming up. J. Control. Release 69, 413–419. oxidation and lipase digestibility. J. Agric. Food Chem. 59, 415–427.
Hogan, J.E., 1989. Hydroxypropylmethylcellulose sustained release technology. Lemarchand, C., Couvreur, P., Vauthier, C., Costantini, D., Gref, R., 2003. Study of
Drug Dev. Ind. Pharma. 15, 975–999. emulsion stabilization by graft copolymers using the optical analyzer Turbiscan.
Hunter, R.J., 1986. Transport properties of suspensions. In: Hunter, R.J. (Ed.), Founda- Int. J. Pharma. 254, 77–82.
tion of Colloid Science, vol. 1. Oxford University Press, New York, pp. 494–536. Leroux, J., Langendorff, V., Schick, G., Vaishnav, V., Mazoyer, J., 2003. Emulsion sta-
Idris, O.H.M., Williams, P.A., Phillips, G.O., 1998. Characterization of gum from Acacia bilizing properties of pectin. Food Hydrocolloids 17, 455–462.
senegal trees of different age and location using multidetection gel permeation Ley, J.P., 2008. Masking bitter taste by molecules. Chemosens. Percept. 1, 58–77.
chromatography. Food Hydrocolloids 12, 379–388. Li, F., Wang, T., He, H.B., Tang, X., 2008. The properties of bufadienolides-loaded
Iglin, P., Avci, G., Silan, C., Ekici, S., Aktas, N., Ayyala, R.S., John, V.T., Sahiner, N., nano-emulsion and submicro-emulsion during lyophilization. Int. J. Pharma.
2010. Colloidal drug carriers from (sub)micron hyaluronic acid hydrogel parti- 349, 291–299.
cles with tunable properties for biomedical applications. Carbohydr. Polymers Li, Y., Hu, M., Xiao, H., Du, Y., Decker, E.A., McClements, D.J., 2010. Controlloing the
82, 997–1003. functional performance of emulsion-based delivery systems using multicompo-
Ivanov, I.B., Danov, K.D., Ananthapadmanabhan, K.P., Lips, A., 2005. Interfacial rhe- nent biopolymer coatings. Eur. J. Pharma. Biopharma. 76, 38–47.
ology of adsorbed layers with surface reaction: on the origin of the dilatational Li, Y., Zheng, J., Xiao, H., McClements, D.J., 2012. Nanoemulsion-based delivery
surface viscosity. Adv. Colloid Interface Sci. 114–115, 61–92. systems for poorly water-soluble bioactive compounds: Influence of formula-
Jansson, P.-E., Kenne, L., Lindberg, B., 1975. Structure of the extracellular polysac- tion parameters on polymethoxyflavone crystallization. Food Hydrocolloids 27,
charide from Xanthomonas campestris. Carbohydr. Res. 45, 275–282. 517–528.
Jenkins, P., Ralston, J., 1998. The adsorption of a polysaccharide at the talc- Liwarska-Bizukojc, E., Miksch, K., Malachowska-Jutsz, A., Kalka, J., 2005. Acute
aqueous solution interface. Colloids Surf. A: Physicochem. Eng. Aspects 139, toxicity and genotoxicity of five selected anionic and non-ionic surfactants.
27–40. Chemosphere 58, 1249–1253.
E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378 377

Lobo, L., 2002. Coalescence during emulsification; 3. Effect of gelatin on rupture and Murillo-Martinez, M.M., Pedroza-Islas, R., Lobato-Calleros, C., Martinez-Ferez, A.,
coalescence. J. Colloid Interface Sci. 254, 165–174. Vernon-Carter, E.J., 2011. Designing W1 /O/W2 double emulsions stabilized
Lourenço, M.J.L., Sampaio, J.P., 2009. Microbial deterioration of gelatin emulsion by protein-polysaccharide complexes for producing edible films: rheological,
photographs: differences of susceptibility between black and white and colour mechanical and water vapour properties. Food Hydrocolloids 25, 577–585.
materials. Int. Biodeteriorat. Biodegrad. 63, 496–502. Musampa, R.M., Alves, M.M., Maia, J.M., 2007. Phase separation, rheology and
Lu, E.-X., Jiang, Z.-Q., Zhang, Q.-Z., Jiang, X.-G., 2003. A water-insoluble drug mono- microstructure of pea protein-kappa-carrageenan mixtures. Food Hydrocolloids
lithic osmotic tablet system utilizing gum arabic as an osmotic, suspending and 21, 92–99.
expanding agent. J. Control. Release 92, 375–382. Myers, D., 1999. Surfaces, Interfaces and Colloids Principles and Applications. Wiley-
Lutz, R., Aserin, A., Wicker, L., Garti, N., 2009. Double emulsions stabilized by a VCH, USA.
charged complex of modified pectin and whey protein isolate. Colloids Surf. Nahringbauer, I., 1995. Dynamic surface tension of aqueous polymer solution, I:
B: Biointerf. 72, 121–127. ethyl(hydroxyethyl)cellulose (BERMOCOLL cst-103). J. Colloid Interface Sci. 176,
Mackie, A.R., Gunning, A.P., Wilde, P.J., Morris, V.J., 2000. Orogenic displacement of 318–328.
protein from the oil/water interface. Langmuir 16, 2242–2247. No, H.K., Meyers, S.P., Prinyawiwatkul, W., Xu, Z., 2007. Application of chitosan for
Mackie, W., Noy, R., Sellen, D.B., 1980. Solution properties of sodium alginates. improvement of quality and shelf life of foods: a review. J. Food Sci. 72, 87–100.
Biopolymers 19, 1839–1860. Norde, W., 2003. Colloids and Interfaces in Life Sciences. Marcel Dekker, New York.
Maldonado-Vaderrama, J., Rodriguez Patino, J.M., 2010. Interfacial rheology O’Regan, J., Mulvihill, D.M., 2010. Sodium caseinate-maltodextrin conjugate stabi-
of proteína-surfactant mixtures. Curr. Opin. Colloid Interface Sci. 15, lized double emulsions: encapsulation and stability. Food Res. Int. 43, 224–231.
271–282. Ochoa-Machiste, E., Buckton, G., 1996. Dynamic surface tension studies of hydro-
Marcuzzo, E., Sensidoni, A., Debeaufort, F., Voilley, A., 2010. Encapsulation of aroma xypropylmethylcellulose film-coating solutions. Int. J. Pharma. 145, 197–201.
compounds in biopolymeric emulsion based edible films to control flavor Ogawa, S., Decker, E.A., McClements, D.J., 2004. Production and characterization
release. Carbohydr. Polymers 80, 984–988. of O/W emulsions containing droplets stabilized by lecithin−chitosan−pectin
Mason, T.G., Wilking, J.N., Meleson, K., Chang, C.B., Graves, S.M., 2006. Nanoemul- mutilayered membranes. J. Agric. Food Chem. 52, 3595–3600.
sions: formation, structure, and physical properties. J. Phys.: Condensed Matter. Oh, E.J., Park, K., Kim, K.S., Kim, J., Yang, J.A., Kong, J.H., Lee, M.Y., 2010. Target specific
18, R635–R666. and long-acting delivery of protein, peptide, and nucleotide therapeutics using
McClements, D.J., 2000. Comments on viscosity enhancement and depletion floccu- hyaluronic acid derivatives. J. Control. Release 141, 2–12.
lation by polysaccharides. Food Hydrocolloids 14, 173–177. Ould Eleya, M.M., Turgeon, S.L., 2000. Rheology of $-carrageenan and "-
McClements, D.J., 2004. Protein-stabilized emulsions. Curr. Opin. Colloid Interface lactoglobulin mixed gels. Food hydrocolloids 14, 29–40.
Sci. 9, 305–313. Padala, S.R., Williams, P.A., Philips, G.O., 2009. Adsorption of gum Arabic, egg white
McClements, D.J., 2011. Edible nanoemulsions: fabrication, properties, and func- protein and their mixtures at the oil-water interface in limonene oil-in-water
tional performance. Soft Matter. 7, 2297–2316. emulsions. J. Agric. Food Chem. 57, 4964–4973.
McClements, D.J., 2012. Nanoemulsions versus microemulsions: terminology, dif- Paraskevopoulou, A., Boskou, D., Kiosseoglou, V., 2004. Stabilization of olive
ferences, and and similarities. Soft Matter. 8, 1719–1729. oil–lemon juice emulsion with polysaccharides. Food Chem. 90, 627–634.
McClements, D.J., Li, Y., 2010. Structured emulsion-based delivery systems: con- Paulsson, M., Dejmek, P., 1992. Surface film pressure of "-lactoglobulin, #-
trolling the digestion and release of lipophilic food components. Adv. Colloid lactalbumin and bovine serum albumin at the air/water interface studied by
Interface Sci. 159, 213–228. Wilhelmy plate and drop volume. J. Colloid Interface Sci. 150, 394–403.
McKinnon, A.A., Rees, D.A., Williamson, F.B., 1969. Coil to double helix transition for Payet, L., Terentjev, E.M., 2008. Emulsification and stabilization mechanisms of O/W
a polysaccharide. Chem. Commun. 701, 702. emulsions in the presence of chitosan. Langmuir 24, 12247–12252.
McNamee, B.F., O’Riordan, E.D., O’Sullivan, M., 1998. Emulsification and microen- Peh, K.K., Wong, C.F., 2000. Application of similarity factor in the development of
capsulation properties of gum arabic. J. Agric. Food Chem. 46, 4551–4555. controlled release diltiazem tablet. Drug Dev. Ind. Pharma. 26, 723–730.
Melton, L.D., Mindt, L., Rees, D.A., Sanderson, G.R., 1976. Covalent structure of the Perez, A.A., Carrara, C.R., Sanchez, C.C., Santiago, L.G., Rodriguez Patino, J.M., 2009.
extracellular polysaccharide from Xanthomonas campestris: evidence from par- Interfacial dynamic properties of whey protein concentrate/polysaccharide mix-
tial hydrolysis studies. Carbohydr. Res. 46, 245–257. tures at neutral pH. Food Hydrocolloids 23, 1253–1262.
Mengual, O., Meunier, G., Cayré, I., Puech, K., Snabre, P., 1999. TURBISCAN MA 2000: Perez, A.A., Sanchez, C.C., Rodriguez Patino, J.M., Pilosof, A.M.R., 2006. Thermody-
multiple light scattering measurements for concentrated emulsion and suspen- namic and dynamic characteristics of hydroxypropylmethylcellulose adsorbed
sion instability analysis. Talanta 50, 445–456. films at the air–water interface. Biomacromolecules 7, 388–393.
Mezdour, S., Lepine, A., Erazo-Majewics, P., Ducept, F., Michon, C., 2008. Oil/water Perez, A.A., Sanchez, C.C., Rodriguez Patino, J.M., Rubiolo, A.C., Santiago, L.G.,
surface rheological properties of hydroxypropyl cellulose (HPC) alone and mixed 2012. Effect of enzymatic hydrolysis and polysaccharide addition on the "-
with lecithin: contribution to emulsion stability. Colloids Surf. A: Physicochem. lactoglobulin adsorption at the air–water interface. J. Food Eng. 109, 712–720.
Eng. Aspects 331, 76–83. Perrechil, F.A., Cunha, R.L., 2010. Oil-in-water emulsions stabilized by sodium
Mikkonen, K.S., Tenkanen, M., Cooke, P., Xu, C., Rita, H., Willför, S., Holmbom, B., caseinate: influence of pH, high-pressure homogenization and locust bean gum
Hicks, K., Yadav, M.P., 2009. Mannans as stabilizers of oil-in-water beverage addition. J. Food Eng. 97, 441–448.
emulsions. LWT – Food Sci. Technol. 42, 849–855. Peter Jr., T., 1975. In: Putnam, F.W. (Ed.), The Plasma Proteins. Academic Press, New
Miller, R., Fainerman, V.B., Makievski, A.V., Krägel, J., Grigoriev, D.O., Kazakov, V.N., York, p. 133.
Sinyachenko, O.V., 2000. Dynamics of protein and mixed protein/surfactant lay- Phaechamud, T., Ritthidej, G.C., 2007. Sustained release from layered matrix system
ers at the water/fluid interface. Adv. Colloid Interface Sci. 86, 39–82. comprising chitosan and xanthan gum. Drug Dev. Ind. Pharma. 33, 595–605.
Mitidieri, F.E., Wagner, J.R., 2002. Coalescence of o/w emulsions stabilized by whey Picton, L., Bataille, I., Muller, G., 2000. Analysis of a complex polysaccharide (gum
and isolate soybean proteins. Influence of thermal denaturation, salt addition arabic) by multi-angle laser light scattering coupled on-line to size exclusion
and competitive interfacial adsorption. Food Res. Int. 35, 547–557. chromatography and flow field flow fractionation. Carbohydr. Polymers 42,
Möbius, D., Miller, R., 1998. Proteins at Liquid Interfaces. Elsevier, Amsterdam. 23–31.
Moe, S.T., Draget, K.I., Skjak-Braek, G., Smidsrød, O., 1995. Alginates. In: Stephen, Pisárčik, M., Bakoš, D., Čeppan, M., 1995. Non-Newtonian properties of hyaluronic
A.M. (Ed.), Food polysaccharides and their applications. US Marcel Decker, New acid aqueous solution. Colloids Surf. A: Physicochem. Eng. Aspects 97, 197–202.
York, pp. 245–286. Pongsawatmanit, R., Harnsilawat, T., McClements, D.J., 2006. Influence of alginate,
Moreau, L., Kim, H.J., Decker, E.A., McClements, D.J., 2003. Production and pH and ultrasound treatment on palm oil-in-water emulsions stabilized by "-
characterization of oil-in-water emulsions containing droplets stabilized by "- lactoglobulin. Colloids Surf. A: Physicochem. Eng. Aspects 287, 59–67.
lactoglobulin-pectin membranes. J. Agric. Food Chem. 51, 6612–6617. Prakash, U.R.T., Thiagarajan, P., 2011. Nanoemulsions for drug delivery through dif-
Morris, E.R., Rees, D.A., Robinson, G., 1980. Cation-specific aggregation of car- ferent routes. Res. Biotechnol. 2, 1–13.
rageenan helices: domain model of polymer gel structure. J. Mol. Biol. 138, Qian, C., Decker, E.A., Xiao, H., McClements, D.J., 2012. Physical and chemical sta-
349–362. bility of "-carotene-enriched nanoemulsions: influence of pH, ionic strength,
Morris, E.R., Rees, D.A., Thom, D., Boyd, J., 1978. Chiroptical and stoechiometric evi- temperature, and emulsifier type. Food Chem. 132, 1221–1229.
dence of a specific, primary dimerisation process in alginate gelation. Carbohydr. Qian, C., McClements, D.J., 2011. Formation of nanoemulsions stabilized by model
Res. 66, 145–154. food-grade emulsifiers using high-pressure homogenization: factors affecting
Moschakis, T., Murray, B.S., Biliaderis, C.G., 2010. Modifications in stability and struc- particle size. Food Hydrocolloids 25, 1000–1008.
ture of whey protein-coated o/w emulsions by interacting chitosan and gum Ramakrishnan, A., Pandit, N., Badgujar, M., Bhaskar, C., Rao, M., 2007. Encapsula-
arabic mixed dispersions. Food Hydrocolloids 24, 8–17. tion of endoglucanase using a biopolymer gum arabic for its controlled release.
Moschakis, T., Murray, B.S., Dickinson, E., 2005. Microstructural evolution of vis- Bioresour. Technol. 98, 368–372.
coelastic emulsions stabilized by sodium caseinate and xanthan gum. J. Colloid Randall, R.C., Phillips, G.O., Williams, P.A., 1988. The role of the proteinaceous com-
Interface Sci. 284, 714–728. ponent on the emulsifying properties of gum arabic. Food Hydrocolloids 2,
Mouécoucou, J., Frémont, S., Villaume, C., Sanchez, C., Méjean, L., 2004. Polysaccha- 131–140.
rides reduce in vitro IgG/IgE-binding of "-lactoglobulin after hydrolysis. Food Randall, R.C., Phillips, G.O., Williams, P.A., 1989. Fractionation and characterization
Chem. 104, 1242–1249. of gum from Acacia senegal. Food Hydrocolloids 3, 65–75.
Muhiddinov, Z., Khalikov, D., Speaker, T., Fassihi, R., 2004. Development and Ravera, F., Loglio, G., Kovalchuk, V.I., 2010. Interfacial dilational rheology by oscil-
characterization of different low methoxy pectin microcapsules by an emulsion- lating bubble/drop methods. Curr. Opin. Colloid Interface Sci. 15, 217–228.
interface reaction technique. J. Microencapsul. 21, 729–741. Razumovsky, L., Damodaran, S., 1999. Surface activity–compressibility relationship
Murata, Y., Miyashita, M., Kofuji, K., Miyamoto, E., Kawashima, S., 2004. Drug release of proteins. Langmuir 15, 1392–1399.
properties of gel bead prepared with pectin and hydrolysate. J. Control. Release Rodriguez, M.S., Albertengo, L.A., Agullo, E., 2002. Emulsification capacity of chitosan.
95, 61–66. Carbohydr. Polymers 48, 271–276.
378 E. Bouyer et al. / International Journal of Pharmaceutics 436 (2012) 359–378

Rooker, W.A., 1927. New uses of fruit pectin. Fruit Prod. J. Am. Vinegar Ind. 7, 11. Sun, C., Gunasekaran, S., Richards, M.P., 2007. Effect of xanthan gum on physico-
Ropers, M.H., Novales, B., Boué, F., Axelos, M.A.V., 2008. Polysaccharide/surfactant chemical properties of whey protein isolate stabilized oil-in-water emulsions.
complexes at the air–water interface – effect of the charge density on interfacial Food Hydrocolloids 21, 555–564.
and foaming behaviors. Langmuir 24, 12849–12857. Surh, J., Gu, Y.S., Decker, E.A., McClements, D.J., 2005. Influence of environmental
Rosilio, V., Boissonnade, M.-M., Zhang, J., Jiang, L., Baszkin, A., 1997. Penetration of stresses on stability of O/W emulsions containing cationic droplets stabilized by
glucose oxidase into organized phospholipid monolayers spread at the solu- SDS-fish gelatin membranes. J. Agric. Food Chem. 53, 4236–4244.
tion/air interface. Langmuir 13, 4669–4675. Surh, J., Decker, E.A., McClements, D.J., 2006. Properties and stability of oil-in-water
Rowe, R.C., 1980. The molecular weight and molecular weight distribution of emulsions stabilized by fish gelatin. Food Hydrocolloids 20, 596–606.
hydroxypropyl methylcellulose used in the film coating of tablets. J. Pharma. Suttiprasit, P., Al-Malah, K., McGuire, J., 1993. On evaluating the emulsifying pro-
Pharmacol. 32, 116–119. perties of protein using conductivity measurements. Food Hydrocolloids 7,
Rowe, R.C., Sheskey, P.J., Owen, S.C., 2006. Handbook of Pharmaceutical Excipients, 241–253.
5th edition. Pharmaceutical Press and the American Pharmacists Association, Tadros, T.F., 1994. Fundamental principles of emulsion rheology and their applica-
London. tions. Colloids Surf. A: Physicochem. Eng. Aspects 91, 39–55.
Ru, Q., Cho, Y., Huang, Q., 2009. Biopolymer-stabilized emulsions on the basis Tadros, T.F., 2009. Emulsion science and technology: a general introduction. In:
of interactions between "-lactoglobulin and %-carrageenan. Front. Chem. Eng. Tadros, T.F. (Ed.), Emulsion Science and Technology. Wiley-VCH Verlag GmbH
China 3, 399–406. and Co. KGaA, Weinheim, pp. 1–56.
Russev, S.C., Arguirov, T.V.1, Gurkov, T.D., 2000. "-Casein adsorption kinetics on Tadros, T.F., Vincent, B., 1983. Emulsion stability. In: Becher, P. (Ed.), Encyclopedia
air–water and oil–water interfaces studied by ellipsometry. Colloids Surf. B: of Emulsion Technology, vol. 1. Marcel Dekker, New York, pp. 129–285.
Biointerf. 19, 89–100. Taherian, A.R., Britten, M., Sabik, H., Fustier, P., 2011. Ability of whey protein isolate
Sanchez, C., Renard, D., Robert, P., Schmitt, C., Lefebvre, J., 2002. Structure and rheo- and/or fish gelatin to inhibit physical separation and lipid oxidation in fish oil-
logical properties of Acacia gum dispersions. Food Hydrocolloids 16, 257–267. in-water beverage emulsion. Food Hydrocolloids 25, 868–878.
Sapei, L., Naqvi, M.A., Rousseau, D., 2012. Stability and release properties of double Tcholakova, S., Denkov, N.D., Ivanov, I.B., Campbell, B., 2006. Coalescence stability of
emulsions for food applications. Food Hydrocolloids 27, 316–323. emulsions containing globular milk proteins. Adv. Colloid Interface Sci. 123–126,
Scherlunda, M., Malmstenb, M., Brodina, A., 1998. Stabilization of a thermoset- 259–293.
ting emulsion system using ionic and nonionic surfactants. Int. J. Pharma. 173, Tcholakova, S., Denkov, N.D., Sidzhakova, D., Ivanov, I.B., Campbell, B., 2005. Effects of
103–116. electrolyte concentration and pH on the coalescence stability of "-lactoglobulin
Schubert, H., Engel, R., 2004. Product and formulation engineering of emulsions. emulsions: experiment and interpretation. Langmuir 21, 4842–4855.
Chem. Eng. Res. Des. 82, 1137–1143. Thanasukarn, P., Pongsawatmanit, R., McClements, D.J., 2006. Utilization of layer-
Schulz, P.C., Rodriguez, M.S., Del Blanco, L.F., Pistonesi, M., Agullo, E., 1998. Emulsi- by-layer interfacial deposition technique to improve freeze-thaw stability of
fication properties of chitosan. Colloid Polymer Sci. 276, 1159–1165. oil-in-water emulsions. Food Res. Int. 39, 721–729.
Serrien, G., Geeraerts, G., Ghosh, L., Joos, P., 1992. Dynamic surface properties Therkelsen, G.H., 1993. Carrageenan. In: Whistler, R.L., BeMiller, J.N. (Eds.), Industrial
of adsorbed protein solutions: BSA, casein and buttermilk. Colloids Surf. 68, Gums. , 3rd ed. Academic Press, San Diego, CA, pp. 145–180.
219–233. Valenta, C., Schultz, K., 2004. Influence of carrageenan on the rheology and skin
Seta, L., Baldino, N., Gabriele, D., Lupi, F.R., de Cindio, B., 2012. The effect of the sur- permeation of microemulsion formulations. J. Control. Release 95, 257–265.
factant type on the rheology of ovalbumin layers at the air/water and oil/water Varum, K.M., Smidsrød, O., 2006. Chitosans. In: Stephen, A.M., Phillips, G.O.,
interfaces. Food Hydrocolloids 29, 247–257. Williams, P.A. (Eds.), Food Polysaccharides and Their Applications. CRC Press,
Sezer, A.D., Akbuğa., 1995. Controlled release of piroxicam from chitosan beads. Int. pp. 497–520.
J. Pharma. 121, 113–116. Vasconcellos, F.C., Goulart, G.A.S., Beppu, M.M., 2011. Production and charac-
Shahidi, F., Abuzaytoun, R., 2005. Chitin, chitosan and co-products: chemistry, pro- terization of chitosan microparticles containing papain for controlled release
duction, applications and health effects. Adv. Food Nutr. Res. 49, 93–135. applications. Powder Technol. 205, 65–70.
Sheppard, S.E., 1929. The function of gelatin in photographic emulsions. J. Frank. Vlachy, N., Touraud, D., Heilmann, J., Kunz, W., 2009. Determining the cytotoxi-
Inst. 208, 559. city of cationic surfactant mixtures on HeLa cells. Colloids Surf. B: Biointerf. 70,
Siew, C.K., Williams, P.A., 2008. Role of protein and ferulic acid in the emulsification 278–280.
properties of sugar beet pectin. J. Agric. Food Chem. 56, 4164–4171. Walstra, P., 1987. Physical principles of emulsion science. In: Blanshard, J., Lillford,
Singh, K.K., 2006a. Xanthan gum. In: Rowe, R.C., Sheskey, P.J., Owen, S.C. (Eds.), Hand- P.J. (Eds.), Food Structure and Behaviour. Academic Press, London, pp. 87–106.
book of pharmaceutical excipients, 5th ed. Pharmaceutical Press and American Ward, A.F.H., Tordai, L., 1946. Time-dependence of bondary tensions of solutions. J.
Pharmacist Association, London, pp. 821–823. Chem. Phys. 14, 453–461.
Singh, K.K., 2006b. Carrageenan. In: Rowe, R.C., Sheskey, P.J., Owen, S.C. (Eds.), Hand- Ward, A.J.I., Regan, L.H., 1980. Pendant drop studies of adsorbed films of BSA. I.
book of Pharmaceutical Excipients. , 5th ed. Pharmaceutical Press and American Interfacial tensions at the Isooctane/water interface. J. Colloid Interface Sci. 78,
Pharmacist Association, London, pp. 124–127. 389–394.
Singh, H., Tamehana, M., Hemar, Y., Munro, P.A., 2003. Interfacial compositions, Weiss, J., Scherze, I., Muschiolok, G., 2005. Polysaccharide gel with multiple emul-
microstructure and stability of oil-in-water emulsions formed with mixtures sion. Food Hydrocolloids 19, 605–615.
of milk proteins and $-carrageenan: 2. Whey protein isolate (WPI). Food Hydro- Wilde, P., Mackie, A., Husband, F., Gunning, P., Morris, V., 2004. Proteins and emul-
colloids 17, 549–561. sifiers at liquid interfaces. Adv. Colloid Interface Sci. 108-109, 63–71.
Singla, A.K., Chawla, M., 2001. Chitosan: some pharmaceutical and biological aspects Wollenweber, C., Makievski, A.V., Miller, R., Daniels, R., 2000. Adsorption of hydroxy-
– an update. J. Pharma. Pharmacol. 53, 1047–1067. propylmethylcellulose at liquid/liquid interface and the effect on emulsion sta-
Sittikijyothin, W., Torres, D., Gonçalves, M.P., 2005. Modelling the rheologi- bility. Colloids Surf. A: Physicochem. Eng. Aspects 172, 91–101.
cal behaviour of galactomannan aqueous solutions. Carbohydr. Polymers 59, Wooster, T.J., Ann Augustin, M., 2006. "-Lactoglobulin-dextran Maillard conjugates:
339–350. their effect on interfacial thickness and emulsion stability. J. Colloid Interface Sci.
Skaugrud, O., 1991. Chitosan – new biopolymer for cosmetics and drugs. Drug Cos- 303, 564–572.
metic Ind. 148, 24–29. Wooster, T.J., Golding, M., Sanguansri, P., 2008. Impact of oil type on nanoemulsion
Smidsrød, O., 1970. Solution properties of alginates. Carbohydr. Polymers 3, formation and Ostwald ripening stability. Langmuir 24, 12758–12765.
359–372. Wu, Y., Cui, W., Eskin, N.A.M., Goff, H.D., 2009. An investigation of four commercial
Solans, C., Izquierdo, P., Nolla, J., Azemar, N., Garcia-Celma, M.J., 2005. Nano- galactomannans on their emulsion and rheological properties. Food Res. Int. 42,
emulsions. Curr. Opin. Colloid Interface Sci. 10, 102–110. 1141–1146.
Solans, C., Pons, R., Kunieda, H., 1997. Overview of basic aspects of microemulsions. Wüstnek, R., Moser, B., Muschiolik, G., 1999. Interfacial dilational behaviour of
In: Solans, C, Kunieda, H (Eds.), Industrial Applications of Microemulsions, vol. adsorbed "-lactoglobulin layers at the different fluid interfaces. Colloids Surf.
66. Marcel Dekker, New York, pp. 1–19. B: Biointerf. 15, 263–273.
Sosa-Herrera, M.G., Berli, C.L.A., Martinez-Padilla, L.P., 2008. Physicochemical Ye, A., 2010. Surface protein composition and concentration of whey protein isolate-
and rheological properties of oil-in-water emulsions prepared with sodium stabilized oil-in-water emulsions: effect of heat treatment. Colloids Surf. B:
caseinate/gellan gum mixtures. Food Hydrocolloids 22, 934–942. Biointerf. 78, 24–29.
Stanley, N.F., 1990. Carrageenans. In: Harris, P. (Ed.), Food Gels. Elsevier Science Yeole, P.G., Galgatte, U.C., Babla, I.B., Nakhat, P.D., 2006. Design and evaluation
Publishers, Barking, UK, pp. 79–119. of Xanthan gum-based sustained release Matrix tablets of Diclofenac sodium.
Su, J., Flanagan, J., Hemar, Y., Singh, H., 2006. Synergistic effects of polyglycerols Indian J. Pharma. Sci. 68, 185–189.
ester of polyricinoleic acid and sodium caseinate on the stabilisation of water- Youssef, M.K., Wang, Q., Cui, S.W., Barbut, S., 2009. Purification and partial physico-
oil-water emulsions. Food Hydrocolloids 20, 261–268. chemical characteristics of protein free fenugreek gums. Food Hydrocolloids 23,
Su, J., Flanagan, J., Singh, H., 2008. Improving encapsulation efficiency and stabilitu 2049–2053.
of water-oil-water emulsions using a modified gum Arabic (Acacia (sen) SUPER Yun, Y.H., Goetz, D.J., Yellen, P., Chen, W., 2004. Hyaluronan microspheres for sus-
GUMTM ). Food Hydrocolloids 22, 112–120. tained gene delivery and site-specific targeting. Biomaterials 25, 147–157.
Sun, C., Gunasekaran, S., 2009. Effects of protein concentration and oil-phase vol- Zelisko, P.M., Flora, K.K., Brennan, J.D., Brook, M.A., 2008. Water-in-silicone
ume fraction on the stability and rheology of menhaden oil-in-water emulsions oil emulsion stabilizing surfactants formed from native albumin and
stabilized by whey protein isolate with xanthan gum. Food Hydrocolloids 23, #, (-triethoxysilylpropyl-polydimethylsiloxane. Biomacromolecules 9,
165–174. 2153–2161.

You might also like