You are on page 1of 47

Critical Reviews in Food Science and Nutrition

ISSN: 1040-8398 (Print) 1549-7852 (Online) Journal homepage: http://www.tandfonline.com/loi/bfsn20

Food-Grade Nanoemulsions: Formulation,


Fabrication, Properties, Performance, Biological
Fate, and Potential Toxicity

David Julian McClements & Jiajia Rao

To cite this article: David Julian McClements & Jiajia Rao (2011) Food-Grade Nanoemulsions:
Formulation, Fabrication, Properties, Performance, Biological Fate, and Potential Toxicity, Critical
Reviews in Food Science and Nutrition, 51:4, 285-330, DOI: 10.1080/10408398.2011.559558

To link to this article: http://dx.doi.org/10.1080/10408398.2011.559558

Published online: 22 Mar 2011.

Submit your article to this journal

Article views: 5336

View related articles

Citing articles: 334 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=bfsn20

Download by: [Australian Catholic University] Date: 30 July 2017, At: 01:41
Critical Reviews in Food Science and Nutrition, 51:285–330 (2011)
Copyright C Taylor and Francis Group, LLC
ISSN: 1040-8398 print / 1549-7852 online
DOI: 10.1080/10408398.2011.559558

Food-Grade Nanoemulsions:
Formulation, Fabrication, Properties,
Performance, Biological Fate, and
Potential Toxicity

DAVID JULIAN McCLEMENTS and JIAJIA RAO


Department of Food Science, University of Massachusetts, Amherst, MA

Nanoemulsions fabricated from food-grade ingredients are being increasingly utilized in the food industry to encapsulate,
protect, and deliver lipophilic functional components, such as biologically-active lipids (e.g., ω-3 fatty acids, conjugated
linoleic acid) and oil-soluble flavors, vitamins, preservatives, and nutraceuticals. The small size of the particles in nanoemul-
sions (r < 100 nm) means that they have a number of potential advantages over conventional emulsions—higher stability
to droplet aggregation and gravitational separation, high optical clarity, ability to modulate product texture, and, increased
bioavailability of lipophilic components. On the other hand, there may also be some risks associated with the oral ingestion
of nanoemulsions, such as their ability to change the biological fate of bioactive components within the gastrointestinal tract
and the potential toxicity of some of the components used in their fabrication. This review article provides an overview of
the current status of nanoemulsion formulation, fabrication, properties, applications, biological fate, and potential toxicity
with emphasis on systems suitable for utilization within the food and beverage industry.

Keywords nanoemulsions, emulsions, delivery systems, encapsulation, toxicology, stability

INTRODUCTION be useful for increasing the bioactivity of some nutraceuticals.


This article provides an overview of the current status of na-
Nanoemulsions have a number of potential advantages over noemulsion formulation, fabrication, properties, applications,
conventional emulsions for particular applications within food biological fate, and potential toxicity, with specific emphasis on
and beverage products. Nanoemulsions usually have better edible systems that are applicable for application within the food
stability to particle aggregation and gravitational separation industry. Only oil-in-water nanoemulsions, which consist of oil
(Tadros et al., 2004). Nanoemulsions contain particles that only droplets dispersed within an aqueous phase, are considered in
scatter light waves weakly, and so they are suitable for incorpo- this review since they currently have the greatest potential for
ration into products that need to be optically clear or only slightly application within commercial products. Nevertheless, water-in-
turbid, such as fortified soft drinks and waters, and some kinds of oil nanoemulsions may also have some interesting applications
soups, sauces, and dips (Mason et al., 2006; Velikov et al., 2008; within certain types of food product.
Wooster et al., 2008). Nanoemulsions can be designed to form
highly viscous or gel-like systems at much lower droplet con-
centrations than conventional emulsions (Sonneville-Aubrun
Nanoemulsions: Terminology
et al., 2004; Tadros et al., 2004; Mason et al., 2006), which
may have interesting applications in some food products. Na-
noemulsions and other colloidal systems have been shown to Initially, we begin by describing and clarifying some of the
increase the bioavailability of certain types of lipophilic sub- terminology commonly used to describe the characteristics of
stances encapsulated within them (Acosta, 2009), which may emulsions, nanoemulsions, and closely-related colloidal disper-
sions (Table 1) (Mason et al., 2006). A conventional emul-
sion, also sometimes known as an emulsion or macro-emulsion,
Address correspondence to D.J. McClements, Department of Food Science,
University of Massachusetts, Amherst, MA 01003. Tel.: 413 545 1019. E-mail: typically has particles with mean radii between 100 nm and
mcclements@foodsci.umass.edu 100 µm. This type of colloidal dispersion is thermodynamically
285
286 D. J. McCLEMENTS AND J. RAO

Table 1 Comparison of thermodynamic stability and physicochemical properties of colloidal dispersions prepared from oil, water and emulsifier. The optical
properties are expressed for typical oil and water systems where there is a significant refractive index contrast, and the particle concentration is appreciable (e.g.,
> 0.1%)

System Droplet Radius Thermodynamic Stability Surface-to-Mass Ratio (m2 /g particles) Optical Properties

Emulsion 100 nm–100 µm Unstable 0.07−70 Turbid/Opaque


Nanoemulsion 10–100 nm Unstable 70−330 Clear/Turbid
Microemulsion 2–100 nm Stable 330−1300 Clear/Turbid

unstable, i.e., the free energy of the separated oil and water nanoemulsions, but much of the material covered will also be
phases is lower than that of the emulsion itself. Consequently, relevant to other types of colloidal dispersions.
conventional emulsions always have a tendency to breakdown Nanoemulsions, like conventional emulsions, can be clas-
over time. Conventional emulsions tend to be optically turbid or sified according to the relative spatial organization of the oil
opaque because the droplets they contain have similar dimen- and water phases. A system consisting of oil droplets dispersed
sions to the wavelength of light (r ≈ λ) and so they scatter light within a water phase is referred to as an oil-in-water (O/W)
strongly (provided the refractive index contrast between the oil nanoemulsion, whereas a system that consists of water droplets
and water phases is not very close to zero). A nanoemulsion, dispersed in an oil phase is referred to as a water-in-oil (W/O)
also referred to as a mini-emulsion, can be considered to be a nanoemulsion. The liquid that makes up the droplets is usually
conventional emulsion that contains very small particles, i.e., referred to as the dispersed phase, whereas the liquid that makes
mean radii between about 10 to 100 nm (Tadros et al., 2004; up the surrounding liquid is referred to as the continuous phase.
Mason et al., 2006). The relatively small particle size compared The particles in an O/W nanoemulsion can be considered to
to the wavelength of light (r << λ) means that nanoemulsions have a core-shell type structure, as represented schematically in
tend to be transparent or only slightly turbid (similar to mi- Fig. 1, with a core made of lipophilic material and a shell made
croemulsions). The very small particle size also means that of surface-active material. As mentioned previously, nanoemul-
they have much better stability to gravitational separation and sions are thermodynamically unfavorable systems because of
aggregation than conventional emulsions (Tadros et al., 2004; the positive free energy associated with creating an interface
Wooster et al., 2008). However, these systems are still thermo- between the oil and water phases due to the hydrophobic effect.
dynamically unstable, and so, will tend to breakdown over time. Consequently, they are metastable systems that tend to break-
In contrast, a micro-emulsion, also sometimes referred to as a down over time due to various physicochemical phenomenon
swollen micelle, is a thermodynamically stable system that typ-
ically contains particles with radii somewhere in the range of
2 to 100 nm. In this case, the free energy of the microemulsion is
lower than that of the phase-separated components from which it
is prepared, so it forms spontaneously. Nevertheless, there may
be kinetic energy barriers between the phase-separated com-
ponents and the microemulsion, which mean that some energy
must be inputted into the system before the microemulsion will
form, e.g., stirring. It is important to note that a microemul-
sion is only thermodynamically stable under a particular set of
environmental conditions (e.g., composition and temperature).
If these conditions are altered, then the microemulsion may no
longer be thermodynamically stable and will be converted to
another kind of system e.g., phase separated, liquid crystalline,
bicontinuous, nanoemulsion, or conventional emulsion. Never-
theless, if the system is brought back to the original conditions
(e.g., composition and temperature), then it should revert back
into a microemulsion, at a rate that depends on any kinetic en-
ergy barriers. Microemulsions are usually transparent because
the particle size is much smaller than the wavelength of light
(r << λ) so that light scattering is weak. From both a scientific Figure 1 Schematic representation of the core-shell structure of particles
and technical point of view, it is important to clearly distin- in nanoemulsions: a shell of amphiphilic material normally covers a core of
lipophilic material. The lipophilic material may contain a variety of different
guish between different kinds of colloidal dispersions since this
non-polar molecules, such as triacylglycerols, diacylglycerols, monoacylglyc-
determines the best approach to optimize their formation, sta- erols, and bioactive components. The amphiphilic shell may contain various
bility, physicochemical properties, and functional performance. surface-active molecules, such as surfactants, phospholipids, proteins, polysac-
In the remainder of this article, the focus will be primarily on charides or minerals.
FOOD-GRADE NANOEMULSIONS 287

Table 2 Impact of various bulk physicochemical properties of oil and aqueous phases on the formation stability and performance of food-grade oil-in-water
nanoemulsions
Property Formation Stability Performance

Viscosity (η) – Efficiency of droplet disruption in high-energy – Creaming rate decreases with – Determines product formation,
– Dispersed (ηD ) methods depends on viscosity ratio (ηD / ηC ) increasing aqueous phase viscosity texture, shelf-life, and release
– Continuous (ηC ) (ηC ) characteristics
Interfacial tension (γ ) – Low γ facilitates droplet disruption in high-energy – Low γ promotes droplet – Affects product formation and
methods; coalescence; shelf-life
– Low γ facilitates droplet formation in – Low γ may lead to poor emulsifier
phase-inversion and spontaneous- emulsification affinity for droplet surfaces
methods
Solubility – Water-solubility of surfactants and solvents affects – High water-solubility of an oil – Affects product shelf-life
droplet formation in solvent displacement and phase promotes Ostwald ripening
spontaneous-emulsification methods
Polarity – – Determines partitioning of – Affects release characteristics &
components between oil and flavor profiles
aqueous phases
Density (ρ) – – Creaming rate increases with – Determines product shelf-life
increasing density contrast
Refractive index (n) – – – Turbidity increases with increasing
refractive index contrast

such as gravitational separation, flocculation, coalescence, facial tension, refractive index, viscosity, density, phase behav-
and/or Ostwald ripening (Dickinson, 1992; Friberg et al., 2004; ior, and chemical stability (Tadros et al., 2004; McClements,
McClements, 2005). The rates at which these degradation 2005; Anton et al., 2007; Wooster et al., 2008; Anton et al.,
processes occur are often considerably different in nanoemul- 2009). These bulk physicochemical characteristics often limit
sions than in conventional emulsions because of differences in the type of homogenization method(s) that can be used to pre-
their particle sizes. Nanoemulsions often have better stability pare a nanoemulsion from a particular oil phase. A number of
to gravitational separation, flocculation, and coalescence than ways in which the bulk physicochemical characteristics of the
conventional emulsions, but are less stable to Ostwald ripening oil phase may influence the formation, stability and performance
(see later). A major concern of food scientists working with of nanoemulsions is summarized in Table 2.
this kind of system is therefore to create nanoemulsions with In the food industry, it is often desirable to prepare na-
sufficiently small particles and then to ensure that they have a noemulsions using triacylglycerol oils due to their low cost,
sufficiently long kinetic stability for commercial applications. abundance, and functional or nutritional attributes, e.g., corn,
The kinetic stability of nanoemulsions can be improved by soybean, sunflower, safflower, olive, flaxseed, algae, or fish oils.
controlling their microstructure (e.g., particle-size distribution), Most of these oils primarily consist of long-chain triacylglyc-
or by incorporating substances known as stabilizers, such as erols (LCT), although medium-chain triacylglycerols (MCT)
emulsifiers, texture modifiers, weighting agents, or ripening and short-chain triacylglycerols (SCT) are being used in some
retarders. (See the following section.) food applications. The formation of nanoemulsions using MCT
and LCT oils is often challenging due to their relatively low
polarity, high interfacial tension, and high viscosity. For ex-
Nanoemulsion Formulation ample, it is difficult to prepare nanoemulsions from these oils
using the phase-inversion temperature (PIT) method because
In this section, we provide an overview of the major compo- of their very high hydrophobicity (Witthayapanyanon et al.,
nents that can be used to formulate food-grade nanoemulsions. 2006). Similarly, it is difficult to prepare nanoemulsions from
these oils using high-pressure homogenization methods because
their high viscosity limits droplet disruption within the homog-
Oil Phase
enizer (Palamakula and Khan, 2004; Wooster et al., 2008; Do
The oil phase used to prepare nanoemulsions can be formu- et al., 2009). In these situations it is often necessary to modify
lated from various non-polar components, including triacylglyc- or optimize conventional homogenization methods to create the
erols, diacylgycerols, monoacylglycerols, free fatty acids, flavor very small droplets needed for nanoemulsion production (see
oils, essential oils, mineral oils, fat substitutes, waxes, weighting later). Once a nanoemulsion has been created using medium- or
agents, oil-soluble vitamins, and various lipophilic nutraceuti- long-chain triacylglycerol oils it is often highly physically sta-
cals (such as carotenoids, curcumin, phytosterols, phytostanols, ble. Conversely, some types of edible oils are highly effective at
and Co-enzyme Q). The formation, stability, and properties of forming nanoemulsions containing small droplets, but are much
nanoemulsions often depend on the bulk physicochemical char- less effective at stabilizing the droplets once they have been pro-
acteristics of the oil phase, e.g., polarity, water-solubility, inter- duced. For example, flavor and essential oils have a relatively
288 D. J. McCLEMENTS AND J. RAO

Figure 2 Nanoemulsions may breakdown through a variety of different physicochemical mechanisms, including gravitational separation, flocculation, coales-
cence, Ostwald ripening, and oiling-off.

high polarity, low interfacial tension, and low viscosity, which to optimize the disperse-to-continuous phase viscosity ratio.
facilitate the formation of very small droplets by both high pres- (See section titled “High Energy Approaches.”) The presence
sure homogenization and PIT methods. However, once this type of cosolvents may also improve the ability of small molecule
of nanoemulsion is formed it may rapidly breakdown because surfactants to form nanoemulsions using phase-inversion meth-
of either Ostwald ripening (due to the high water-solubility of ods. The impact of specialized kinds of molecules, referred to
the oil) or coalescence (due to the low interfacial tension of the as “stabilizers,” which might be added to either aqueous or oil
oil). Again, alternative strategies usually have to be developed phases, is discussed in the following section.
to improve the long-term stability of nanoemulsions created
using these oils, e.g., by adding ripening inhibitors to prevent
Ostwald ripening or by changing the emulsifier type to prevent Stabilizers
coalescence (see later).
If only an oil phase and an aqueous phase are homogenized
together, the system will normally rapidly breakdown through
a variety of different mechanisms, including droplet floccula-
Aqueous Phase
tion, coalescence, Ostwald ripening, and gravitational separa-
The aqueous phase used to prepare a nanoemulsion typically tion (Fig. 2). For this reason, it is often necessary to add various
consists primarily of water, but it may also contain a variety of kinds of stabilizers to nanoemulsions to improve their long-term
other polar components, including co-solvents (such as simple stability.
alcohols and polyols), carbohydrates, proteins, minerals, acids, Emulsifiers/Co-emulsifiers. The selection of an appropriate
and bases. The type and concentration of these components de- emulsifier (or combination of emulsifiers) is one of the most
termines the polarity, interfacial tension, refractive index, rheol- important factors to consider for the proper design of a
ogy, density, phase behavior, pH, and ionic strength of the aque- nanoemulsion. An emulsifier is a surface-active molecule
ous phase, which in turn will impact the formation, stability, that is capable of adsorbing to droplet surfaces, facilitating
and physicochemical properties of the nanoemulsion produced droplet disruption, and protecting droplets against aggregation
(Table 2). Careful control of the aqueous phase composition can (McClements, 2005; Kralova et al., 2009). In high-energy ap-
be used to optimize the formation or improve the stability of na- proaches, the emulsifier facilitates droplet disruption within the
noemulsions. The formation of small droplets in nanoemulsions homogenizer by lowering the interfacial tension, which favors
using high pressure-homogenization methods can often be facil- the production of small droplets. In low-energy approaches,
itated by adding water-soluble cosolvents to the aqueous phase the emulsifier facilitates the spontaneous formation of small
FOOD-GRADE NANOEMULSIONS 289

droplets due to its ability to produce very low interfacial tensions ing a single emulsifier. For example, employing a lipophilic and
under certain environmental and solution conditions (see later). hydrophilic surfactant in conjunction can facilitate the forma-
The stability of a nanoemulsion to environmental stresses such tion of small particles using both low-energy and high-energy
as pH, ionic strength, heating, cooling, or long-term storage is approaches. On the other hand, using mixed-emulsifier systems
often predominantly determined by the kind of emulsifier used. can often improve the stability of nanoemulsions to particle ag-
In the food industry, the most important types of emulsifiers gregation after formation. (See section titled “Finishing Tech-
are small molecule surfactants, phospholipids, proteins, and niques.”)
polysaccharides. The type of emulsifier used has a major impact Some methods of forming nanoemulsions using low-energy
on the type of homogenization approach that can be used to form methods require the presence of cosurfactants (such as short-
a nanoemulsion. Many small molecule surfactants are highly and medium-chain alcohols) or cosolvents (such as polyols like
effective at producing nanoemulsions using both high- and propylene glycol, glycerol, and sorbitol) (Yaghmur et al., 2002;
low-energy methods. On the other hand, proteins and polysac- Flanagan et al., 2006). Cosurfactants are amphiphilic molecules
charides are not typically suitable for producing nanoemulsions that are surface active (often having a hydrocarbon chain and a
using low-energy methods, and are not usually as effective as hydroxyl group), but that are not good at stabilizing emulsions
surfactants at forming nanoemulsions with small droplet sizes themselves due to the small size of the polar head group. A num-
using high-energy methods. On the other hand, proteins and ber of physicochemical mechanisms have been proposed to ac-
polysaccharides have the advantage of being natural ingredients count for the ability of co-surfactants to facilitate nanoemulsion
that are often considered more “label-friendly,” and therefore formation— (1) Fluidization of the interface; (2) Optimization
there is great interest in producing nanoemulsions with them. of the disperse-to-continuous phase viscosity ratio; (3) Reduc-
It is often convenient to classify small molecule surfactants tion of the electrical repulsion between the head groups of ionic
according to their electrical characteristics—ionic, non-ionic, surfactants at an interface by acting as spacers; and, (4) Induc-
and zwitterionic (McClements, 2005): tion of the appropriate interfacial curvature (Gradzielski, 1998;
Garti et al., 2001; Shafiq-un-Nabi et al., 2007). Cosolvents are
• Ionic surfactants: Most food-grade ionic surfactants are neg- highly polar molecules that are not particularly surface active
atively charged (such as CITREM, DATEM, and SLS), but themselves, but that may alter the physicochemical properties of
at least one positively charged surfactant is available for cer- the emulsifier molecules, such as their surface-activity, oil-water
tain applications (i.e., lauric arginate). Ionic surfactants may partition coefficient, and the ability to form colloidal structures.
be used to form nanoemulsions by various low-energy and Cosolvents may do this by altering the structure and interac-
high-energy approaches. Their utilization may be limited in tions of the molecules in the aqueous phase, which alters the
products where high-surfactant levels are required because magnitude of any hydrophobic interactions.
they tend to cause irritation (Sol et al., 2006; Solè et al., Texture Modifier. A texture modifier is a substance that thick-
2006). ens or gels the aqueous phase (Imeson, 2010). Texture modi-
• Non-ionic surfactants: Non-ionic surfactants have been fiers are often incorporated into commercial products to improve
widely used to form nanoemulsions due to their low toxicity, emulsion stability by retarding droplet movement, but they may
lack of irritability, and the capacity to easily form nanoemul- also be used to provide desirable textural characteristics, such as
sions by both high-energy and low-energy approaches. Exam- “creaminess,” “richness,” “thickness,” or gel strength. Most tex-
ples include sugar ester surfactants (e.g., sorbitan monooleate, ture modifiers used in the food industry are biopolymers, such
sucrose monopalmitate), polyoxyethylene ether (POE) sur- as proteins (e.g., egg, milk, vegetable proteins) or polysaccha-
factants (e.g., Brij 97), and ethoxylated sorbitan esters (e.g., rides (e.g., starch, pectin, alginate, carrageenan, xanthan, guar
Tweens and Spans) (Chiu, 2006; Liu et al., 2006; Jafari et al., gum).
2007; Henry et al., 2009). Weighting Agent. A weighting agent is a substance that is
• Zwitterionic surfactants: Zwitterionic surfactants have two usually added to the oil droplets in O/W emulsions to match
or more oppositely charged ionizable groups on the same their density to that of the surrounding aqueous phase (Mc-
molecule. Consequently, they can have a net negative, neutral, Clements, 2005). Its purpose is to reduce the driving force for
or positive charge depending on solution pH. Phospholipids gravitational separation (ρ) and therefore prevent or retard
are common zwitterionic surfactants that have GRAS status, creaming or sedimentation. Commonly used weighting agents
which permits their use in food, e.g., lecithin. However, many in the food and beverage industries include brominated veg-
natural phospholipids are not particularly good at either form- etable oil (BVO), ester gum, damar gum, and sucrose-acetate
ing or stabilizing nanoemulsions when used in isolation, but isobutyrate (SAIB) (McClements, 2005). The density of the
they may be effective when used in combination with co- particles in nanoemulsions may also be matched to that of the
surfactants (Trotta et al., 1996; de Morais et al., 2006; Hoeller surrounding aqueous phase by coating them with thick dense
et al., 2009). layers, or by partially crystallizing the lipid core. (See section
titled “Designing Functional Nanoemulsion Particles.”)
The formation and stability of nanoemulsions can often be Ripening Retarder. The concept of a ripening retarder is
improved by using combinations of emulsifiers, rather than us- relatively new in the food industry, but they are particularly
290 D. J. McCLEMENTS AND J. RAO

important for stabilizing certain kinds of food-grade nanoemul- ical devices that are capable of generating extremely intense dis-
sions, e.g., those containing oils susceptible to Ostwald ripening ruptive forces are capable of producing the tiny droplets required
(OR), such as flavor oils, essential oils, and short-chain triacyl- to form nanoemulsions (Fig. 3), i.e., high-pressure valve homog-
glycerols. A ripening retarder is a highly hydrophobic material enizers, microfluidizers, and ultrasonic devices (Tadros et al.,
that is incorporated into oil droplets to retard or inhibit Ostwald 2004; Leong et al., 2009). The reason that such intense energies
ripening. OR is the process where large droplets grow at the are needed is that the disruptive forces generated by the homog-
expense of smaller droplets due to diffusion of oil molecules enizer must exceed the restoring forces holding the droplets into
through the intervening aqueous phase. (See section titled spherical shapes (Walstra, 1993; Schubert and Behrend, 2003;
“Stability.”) Typically, a ripening retarder is a hydrophobic Schubert and Engel, 2004). These restorative forces are deter-
component that is soluble in the oil phase but highly insoluble in mined by the Laplace Pressure: P = γ /2r,which increases
water, e.g., a long-chain triglyceride, mineral oil, or ester gum with decreasing droplet radius (r) and increasing interfacial
(Kabalnov et al., 1987; Sonneville-Aubrun et al., 2004). A ripen- tension (γ ). Thus, as the droplet radius becomes increasingly
ing retarder works by generating an entropy of mixing effect smaller within a homogenizer, it becomes increasingly difficult
that counteracts the OR effect. (See section titled “Stability.”) to break them up further. The smallest droplet size that can be
produced by a particular high-energy device depends on the ho-
mogenizer design (e.g., flow and force profiles), homogenizer
NANOEMULSION FORMATION operating conditions (e.g., energy intensity, duration), environ-
mental conditions (e.g., temperature), sample composition (e.g.,
There have been considerable advances in our understanding oil type, emulsifier type, concentrations), and the physicochemi-
of the methods that can be used to create nanoemulsions in the cal properties of the component phases (e.g., interfacial tension,
past few years. Nanoemulsions can be fabricated using a num- viscosity) (Kentish et al., 2006; Wooster et al., 2008). Previous
ber of different approaches, but these can usually be broadly studies have shown that the droplet size tends to decrease when
categorized as either high-energy or low-energy approaches de- the energy intensity or duration increases, the interfacial ten-
pending on the underlying principle (Tadros et al., 2004; Pouton sion decreases, the emulsifier adsorption rate increases, and the
and Porter, 2006; Acosta, 2009; Anton et al., 2009; Leong et al., disperse-to-continuous phase viscosity ratio falls within a cer-
2009). tain range (0.05 < ηD /ηC < 5) (Walstra, 1993; Walstra, 2003;
High-energy approaches utilize mechanical devices (“ho- Tadros et al., 2004). The extent of the ηD /ηC range where small
mogenizers”) capable of generating intense disruptive forces droplets can be produced depends on the nature of the disrup-
that are capable of disrupting and intermingling the oil and tive forces generated by the particular homogenizer used, i.e.,
aqueous phases into tiny oil droplets, e.g., high pressure valve simple shear versus extensional flow.
homogenizers, microfluidizers, and sonication methods (Gutier- High-energy approaches are one of the most versatile means
rez et al., 2008; Velikov et al., 2008; Wooster et al., 2008; Leong of producing food-grade nanoemulsions because they can be
et al., 2009). At present, high-energy approaches are the most used with a wide variety of different oil and emulsifier types.
common method used to prepare nanoemulsions in industrial Provided that the homogenization conditions are suitably op-
food operations because their utilization in the production of timized, nanoemulsions can be produced using triacylglycerol
conventional emulsions is already well-established, they are ca- oils, flavor oils, and essentials oils as the oil phase, and proteins,
pable of large-scale production, and they can be used to prepare polysaccharides, phospholipids, and surfactants as emulsifiers.
nanoemulsions from a variety of different starting materials. Even so, the size of the particles produced depends strongly on
Low-energy approaches rely on the spontaneous formation of the characteristics of the oil and emulsifier used (see below).
tiny oil droplets within mixed oil-water-emulsifier systems when For example, it is usually easier to produce very small droplets
the solution or environmental conditions are altered, e.g., phase- when the oil phase has a low viscosity and/or interfacial tension
inversion and spontaneous-emulsification methods (Bouchemal (e.g., flavor oils, essential oils, alkanes) than when it has a high
et al., 2004; Tadros et al., 2004; Freitas et al., 2005; Chu et al., viscosity and/or interfacial tension (e.g., MCT or LCT).
2007a; Anton et al., 2008; Yin et al., 2009). The minimum
particle size that can be produced using each approach depends
High Pressure Valve Homogenizer
on many different factors, which are highlighted in the sections
below. High pressure valve homogenizers are probably the most
common method of producing conventional emulsions with
small droplet sizes in the food industry (Schubert and Behrend,
High-Energy Approaches 2003; Schubert and Engel, 2004). They are more effective at re-
ducing the size of the droplets in pre-existing coarse emulsions,
In general, the particle size produced by high-energy ap- than at creating emulsions directly from two separate liquids. A
proaches is governed by a balance between two opposing pro- coarse emulsion is usually produced using a high shear mixer
cesses occurring within the homogenizer—droplet disruption and is then fed directly into the inlet of the high pressure valve
and droplet coalescence (Jafari et al., 2008). Only those mechan- homogenizer. The homogenizer has a pump that pulls the coarse
FOOD-GRADE NANOEMULSIONS 291

Figure 3 Schematic representation of various mechanical devices that can be used to produce food-grade nanoemulsions using a high-energy approach: high
pressure valve homogenizer; microfluidizer: ultrasonic jet homogenizer; ultrasonic probe homogenizer.

emulsion into a chamber on its backstroke and then forces it less than 100 nm in radius (e.g., high emulsifier levels, low in-
through a narrow valve at the end of the chamber on its forward terfacial tensions, and appropriate viscosity ratios). A number
stroke (Fig. 3). As the coarse emulsion passes through the valve of recent studies that have used high-pressure homogenization
it experiences a combination of intense disruptive forces that to produce nanoemulsions that could potentially be used in the
cause the larger droplets to be broken down into smaller ones. food industry are summarized in Table 3.
Different nozzle designs are available to increase the efficiency
of droplet disruption. The droplet size produced using a high
pressure valve homogenizer usually decreases as the number Microfluidizers
of passes and/or the homogenization pressure increases. It also A microfluidizer is somewhat similar in design to a high-
depends on the viscosity ratio of the two phases (usually oil and pressure homogenizer in that it involves using high pressures
water) being homogenized. Small droplets can only usually be to force an emulsion pre-mix through a narrow orifice to fa-
produced when the disperse-to-continuous phase viscosity ratio cilitate droplet disruption. However, the design of the channels
falls within a certain range (0.05 < ηD /ηC < 5) (Walstra, 1993; through which the emulsion is made to flow within the homog-
Walstra, 2003; Tadros et al., 2004). Finally, it is important to have enizer is different (Fig. 3). Microfluidizers have traditionally
sufficient emulsifier present to cover all the new droplet surfaces been used in the pharmaceutical industry to produce emulsion-
formed during homogenization, and to use an emulsifier that can based products, but they have also been used in the food and
rapidly adsorb to the droplet surfaces to prevent re-coalescence beverage industries to produce flavor emulsions, nutraceutical
(Jafari et al., 2008). Usually, there is a linear relationship be- emulsions, and homogenized milk. The microfluidizer works on
tween the logarithm of the homogenization pressure (P ) and the the principle of dividing an emulsion flowing through a chan-
logarithm of the droplet diameter (d): log d ∝ log P. The con- nel into two streams, passing each stream through a separate
stant of proportionality depends on the homogenizer—for large fine channel, and then directing the two streams at each other
homogenizers and low-viscosity fluids d ∝ P −0.6 ; for large ho- in an interaction chamber. Intense disruptive forces are gener-
mogenizers and high-viscosity fluids d ∝ P −0.75 ; and for small ated within the interaction chamber when the two fast-moving
(bench-top) homogenizers d ∝ P −1.0 (McClements, 2005). To streams of emulsion impinge upon each other, leading to highly
reduce the droplet size to the level required in nanoemulsions it efficient droplet disruption.
is usually necessary to operate at extremely high pressures and A number of studies have examined the potential application
to use multiple passes through the homogenizer. Even then, it of microfluidizers for the production of model food nanoemul-
is only possible under certain circumstances to obtain droplets sions, such as beverage and dairy emulsions (Dalgleish et al.,
292 D. J. McCLEMENTS AND J. RAO

Table 3 An overview of recent research articles on O/W nanoemulsion formation using high pressure valve homogenizers

Oil Phase Bioactive component Surfactant/Co-surfactant Particle Diameter Reference

MCT β−carotene Tween 20,40,60,80 132–148 nm Yuan et al., 2008


MCT β−carotene Tween 20, decaglycerol monolaurate, whey 115–178 nm Mao et al., 2009
protein isolate, modified starch
Oleoresin Capsicum/ethyl Poly-caprolactone Pluronic F68 320–460 nm Choi et al., 2009
acetate
Olive oil Pluronic F68, Phospholipid 239–274 nm Wulff-Perez et al., 2010
MCT Curcumin Tween 20 160 nm Wang et al., 2008

1997; Henry et al., 2010; Klein et al., 2010). These studies have hand, we found little change in mean droplet size with viscosity
shown that small droplets can be produced provided that con- ratio when a globular protein was used as an emulsifier, which
ditions are optimized to facilitate droplet disruption and inhibit again may be due to the relatively slow adsorption of the pro-
droplet coalescence. A number of recent studies carried out us- tein and its ability to form a coating that inhibits further droplet
ing a microfluidizer to produce nanoemulsions are highlighted disruption.
in Table 4. Like high pressure valve homogenizers, the droplet
size produced tends to decrease with increasing homogenization
pressure, number of passes, and emulsifier concentration. In ad- Ultrasonic Homogenizers
dition, the disperse-to-continuous phase viscosity ratio should Ultrasonic homogenizers utilize high-intensity ultrasonic
be within a certain range to facilitate the formation of small waves to create the intense disruptive forces necessary to
droplets (Wooster et al., 2008). As with the high pressure valve breakup up oil and water phases into very small droplets (Ken-
homogenizer, there is usually a linear relationship between the tish et al., 2006; Lin and Chen, 2008; Leong et al., 2009). The
logarithm of homogenization pressure and the logarithm of the energy input is provided using sonicator probes that contain
droplet diameter: log d ∝ log P. Recently, our laboratory per- piezoelectric quartz crystals that expand and contract in response
formed experiments to elucidate the major factors that influence to an alternating electrical voltage (Fig. 3). The tip of the son-
the formation of nanoemulsions using a microfluidizer. The log- icator probe is placed within the liquids being homogenized,
arithm of the mean droplet diameter decreased linearly as the where it generates intense mechanical vibrations that lead to
logarithm of the homogenization pressure increased for both an cavitation effects, i.e., the formation, growth, and collapse of
ionic surfactant (SDS) and a globular protein (β-lactoglobulin) small bubbles in the liquid. The collapse of the micro-bubbles
(Fig. 4). However, the slope of the log(d) vs. log(P ) relationship formed by cavitation generates intense disruptive forces in the
was appreciably steeper for the surfactant (-0.57) than for the immediate vicinity of the sonicator probe that leads to droplet
protein (-0.29), which was attributed to the fact that the protein disruption. For efficient and uniform homogenization it is im-
may adsorb more slowly to the droplet surfaces, and that it may portant to ensure that the emulsion spends sufficient time within
form a viscoelastic coating that inhibits further droplet breakup. the region where droplet disruption occurs. Recently, batch and
The dependence of the mean droplet diameter on the viscosity flow-through ultrasonic homogenizers have been developed that
ratio (ηD /ηC ) was also examined by preparing emulsions us- have special cells designed to optimize the efficiency of droplet
ing different oil-phase compositions (corn oil : octadecane) and disruption (Leong et al., 2009). A number of recent studies car-
aqueous-phase compositions (water : glycerol) (Fig. 5). For the ried out using sonication to produce nanoemulsions are high-
ionic surfactant there was a distinct decrease in mean droplet lighted in Table 5. Experiments have shown that the size of the
diameter with decreasing viscosity ratio, which suggested that droplets produced using an ultrasonic device tends to decrease
droplet disruption within the homogenizer became easier as the as the intensity of the ultrasonic waves is increased, or the resi-
viscosity of the two phases became more similar. On the other dence time in the disruption zone is increased (Abismail et al.,

Table 4 An overview of recent research articles on O/W nanoemulsion formation using a microfluidizer

Oil Phase Bioactive component Surfactant/Co-surfactant∗ Particle Diameter Reference

WPC, WPH, SC, modified starch 150–600 nm Jafari et al., 2007


Peanut oil Tween 80, SDS 120 nm Wooster et al., 2008
Hexadecane Tween 80, SDS 80 nm Wooster et al., 2008
Corn oil/Ethyl acetate WPI 174 nm Lee and McClements, 2010
MCT CO Q10 Nonionic surfactant, soybean lecithin, glycerol 60 nm Hatanaka et al., 2008
Corn oil/Octadecane SDS, Tween20, beta-lactoglobulin 60–150 nm Qian and McClements, 2010
MCT α−Tocopherol Glycerol, Decaglyceryl monooleate Soybean 80–400 nm Hatanaka et al., 2010
lecithin
∗ Here the acronyms WPC, WPH, and SC mean whey protein concentrate, whey protein hydrolysate, and sodium caseinate, respectively.
FOOD-GRADE NANOEMULSIONS 293

Figure 5 Influence of the droplet to continuous phase viscosity ratio on the


Figure 4 Influence of homogenization pressure on the mean particle diameter mean particle diameter of droplets produced using a microfluidizer using either a
of droplets produced using a microfluidizer using either a protein (BLG) or protein (BLG) or surfactant (SDS) as emulsifier (Qian and McClements, 2010).
surfactant (SDS) as emulsifier (Qian and McClements, 2010).

noemulsions are based on the low-energy approach, including


1999; Maa and Tsu, 1999). The homogenization efficiency also spontaneous emulsification- and phase-inversion methods (Fer-
depends on the type and amount of emulsifier present, and the nandez et al., 2004; Pouton and Porter, 2006; Anton et al., 2008;
viscosity of the oil and aqueous phases (Maa and Tsu, 1999; 2009; Maestro et al., 2008).
Jafari et al., 2006; Kentish et al., 2006; Leong et al., 2009). Low-energy approaches are often more effective at producing
All of these parameters must therefore be optimized to produce small droplet sizes than high- energy approaches, but they are
nanoemulsions containing small droplets. There is also some often more limited in the types of oils and emulsifiers that can
concern that the high local intensities involved in sonication be used. For example, it is currently not possible to use proteins
methods may lead to protein denaturation, polysaccharide de- or polysaccharides as emulsifiers in most of the low-energy
polymerization, or lipid oxidation during homogenization. approaches used to form nanoemulsions. Instead, it is often
necessary to utilize relatively high concentrations of synthetic
surfactants to form nanoemulsions by these approaches, which
Low-Energy Approaches may limit their use for many food applications.

The formation of emulsions using low-energy approaches


Spontaneous Emulsification
relies on the spontaneous formation of oil droplets in oil-water-
emulsifier mixtures when either their composition or their en- In this method an emulsion or nanoemulsion is spontaneously
vironment is altered (Bouchemal et al., 2004; Tadros et al., formed when two liquids are mixed together (Miller, 1988; Pou-
2004; Freitas et al., 2005; Chu et al., 2007a; Anton et al., 2008; ton and Porter, 2006; Anton et al., 2009). Systems prepared us-
2009; Yin et al., 2009). A number of methods for preparing na- ing this method are usually referred to as either self-emulsifying

Table 5 An overview of research articles on O/W nanoemulsions using Ultrasonic Homogenizer

Oil phase Surfactant/Co-surfactant Particle Diameter Reference

Medium-chain triglyceride Tween 80/ethanol 20–60 nm Amani et al., 2010


Sunflower oil, Canola oil Tween 80, Span 80, SDS 40 nm Leong et al., 2009
Flax seed oil Tween 40 150 nm Kentish et al., 2008
Olive oil, Sesame oil, Soybean oil Pluronic F68 < 500 nm Wulff-Perez et al., 2009
Soybean oil Phosphatidylcholine / Sodium palmitate, sucrose palmitate 58 nm Takegami, et al., 2008
294 D. J. McCLEMENTS AND J. RAO

Table 6 Some recent research articles on O/W nanoemulsions formed using the Spontaneous Emulsification method that may be suitable for food applications

Oil Phase Bioactive component Surfactant/Co-surfactant Particle Diameter Reference

MCT α-Tocopherol Span 80, Span85, Tween 20, Tween 80/Acetone, Ethanol, Ethyl acetate 100–600 nm Bouchemal et al., 2004
MCT Castor oil Soybean lecithin, Polyoxyl 35 castor oil 150–200 nm Kelmann et al., 2007
MCT Oligothymidylate Egg lecithin/glycerol 225 nm Martini et al., 2008
Palm oil Vitamin E Tween 80, Plurionic F-68 94 nm Teo et al., 2010

drug-delivery systems (SEDDS) or self-nano-emulsifying drug- of droplets (Fig. 6). The size of the droplets produced can be
delivery systems (SNEDDS) in the pharmaceutical industry. A controlled by varying the compositions of the two initial phases,
number of recent studies using the spontaneous emulsification as well as the mixing conditions.
approach are highlighted in Table 6. Practically, this method can Previous studies using the spontaneous emulsification
be carried out in a number of different ways—the compositions method have used one or more synthetic surfactants. To the
of the two phases can be varied; the environmental conditions knowledge of the authors, there are currently no studies where
can be varied (e.g., temperature, pH, and ionic strength); and/or, proteins or polysaccharides have been used as an emulsifier in
the mixing conditions can be varied (e.g., stirring speed, rate of the spontaneous emulsification approach. This may be because
addition, and order of addition). For example, an organic phase of physicochemical limitations associated with the droplet-
consisting of a hydrophobic oil, a hydrophilic surfactant, and a formation mechanism in this method, i.e., the need for small oil
water-miscible organic solvent may be slowly added to water droplets to spontaneously form and bud-off from the oil-water
(Anton et al., 2009). Alternatively, water may be added to an or- interface (Fig. 6). Interfaces covered with biopolymer emulsi-
ganic phase containing hydrophobic oil, water-miscible organic fiers tend to have higher interfacial tensions and more elastic
solvent, and surfactant (Sonneville-Aubrun et al., 2009). The properties than those formed with small molecule surfactants.
water-miscible organic solvent is typically ethanol or acetone, We recently compared the spontaneous emulsification
but in some cases it is possible to replace it with a water-miscible method of producing nanoemulsions with a high energy method.
surfactant instead (Anton et al., 2008; 2009). The surfactant-oil-water system used consisted of 15.4 wt%
A number of physicochemical mechanisms have been pro- non-ionic surfactant, 23.1 wt% MCT, and 61.5% water, with the
posed to account for spontaneous emulsification (Horn and surfactant containing a 50:50 mixture of a hydrophilic (Tween
Rieger, 2001). Consider a situation where two phases are 80) and lipophilic (Tween 85) surfactant. Nanoemulsions were
brought into contact with each other. Each phase contains some prepared using two different methods: (i) High-energy method:
components that are immiscible with the other phase (e.g., one the surfactant, oil and water were blended together, and then
phase contains oil and the other phase contains water), and one passed through a microfluidizer; (ii) Low-energy method: The
of the phases contains another component that is partially misci- oil and surfactant were mixed together, and then simply added
ble with both phases (e.g., an amphiphilic alcohol or surfactant). to the water using slow stirring.
When the two phases are brought into contact, some of the com- The microfluidization method produced droplets with
ponent that is partially miscible with both phases will move from a diameter of about 110 nm, whereas the spontaneous-
its original location in one phase into the other phase. As this emulsification method could produce droplets with diameters
component moves it will cause an increase in oil-water inter- around 140 nm. This simple experiment demonstrated that
facial area, interfacial turbulence, and spontaneous formation nanoemulsions could be produced using the spontaneous

Figure 6 Schematic representation of proposed mechanism for spontaneous emulsification: oil droplets are spontaneously formed when an oil phase containing
a water-dispersible substance is mixed with an aqueous phase. The underlying mechanism is the movement of the water-dispersible substance from the oil phase
to the water phase (red arrows), leading to interfacial turbulence and spontaneous oil droplet formation.
FOOD-GRADE NANOEMULSIONS 295

Figure 7 Schematic diagram of the temperature-dependence of the spontaneous curvature of surfactant monolayers and their influence on emulsion properties.

emulsification method, provided that the system composition temperature (Fig. 7). The molecular geometry of a surfactant
was optimized, i.e., surfactant, oil, and water contents. can be described by a packing parameter, p (Israelachvili,
1992):

aT
Phase-Inversion Methods p=
aH
A number of methods have been developed to formulate
nanoemulsions that depend on inducing a phase inversion in where, aT and aH are the cross-sectional areas of the lipophilic
emulsions from a W/O to O/W form (or vice versa), e.g., tail-group and hydrophilic head-group respectively. Due to the
phase-inversion temperature (PIT), phase-inversion composi- hydrophobic effect, the surfactant molecules tend to sponta-
tion (PIC), and emulsion-inversion point (EIP) methods. We be- neously associate with each other in water and form monolayers
gin by describing the principle underlying each of these meth- that have an optimum curvature that allows the most efficient
ods, and then we conclude by highlighting a recent general packing of the molecules (Israelachvili, 1992). The optimum
conceptual approach that has been developed to rationalize the curvature of a surfactant monolayer depends on the packing pa-
formation of emulsions based on the phase-inversion methods. rameter of the surfactant—for p < 1, the optimum curvature
Phase-Inversion Temperature Methods. The phase- is convex, for p = 1, monolayers with zero curvature are pre-
inversion temperature method (PIT) relies on changes in the ferred, and, for p > 1 the optimum curvature is concave (Fig.
optimum curvature (molecular geometry) or relative solubility 7). In an emulsion or nanoemulsion, the surfactant monolayer
of non-ionic surfactants with changing temperature (Anton et is present at an oil-water interface. The type of emulsion that
al., 2007; 2009; Gutierrez et al., 2008). Nanoemulsions can a particular surfactant tends to favor depends on its molecular
be spontaneously formed using the PIT method by varying geometry. Surfactants with p < 1 tend to favor the formation
the temperature-time profile of certain mixtures of oil, water, of O/W emulsions, those with p > 1 favor W/O emulsions,
and non-ionic surfactant. This type of phase inversion usually and those with p = 1 do not favor either O/W or W/O sys-
involves the controlled transformation of an emulsion from one tems and instead lead to the formation of liquid crystalline or
type to another (e.g., W/O to O/W or vice versa) through an bicontinuous systems (Fig. 7). Changes in the relative solubil-
intermediate liquid crystalline or biocontinuous microemulsion ity of surfactants in oil and water phases with temperature are
phase. The driving force for this type of phase inversion is also needed to understand the physicochemical properties of
changes in the physicochemical properties of the surfactant with non-ionic surfactants (Anton et al., 2007; Anton et al., 2009).
296 D. J. McCLEMENTS AND J. RAO

At low temperatures, the head group of a non-ionic surfactant


is highly hydrated and so it tends to be more soluble in water.
As the temperature is raised, the head group becomes progres-
sively dehydrated and the solubility of the surfactant in water
decreases. At a particular temperature (≈ PIT), the solubility of
the surfactant in the oil and water phases is approximately equal.
At higher temperatures, the surfactant becomes more soluble in
the oil phase than in the water phase.
The temperature dependence of the packing parameter or sol-
ubility of non-ionic surfactants accounts for the ability to form
nanoemulsions using the PIT method (Fig. 7). The temperature
at which an oil-water-surfactant system changes from an O/W
emulsion to a W/O emulsion is known as the phase inversion
temperature or PIT. At temperatures well below the PIT (≈ T <
PIT-30◦ C), the surfactant is soluble in water, its packing param-
eter is appreciably less than unity (p < 1), and it favors the
formation of O/W emulsions. As the temperature is raised, the
hydrophilic head-groups of the non-ionic surfactant molecules
become progressively dehydrated, which causes its water sol-
ubility to decrease, and its packing parameter to tend towards
unity. At the PIT the packing parameter equals unity (p =
1), and the emulsion breaks down because the droplets have
an ultralow interfacial tension and therefore readily coalesce
with each other. The resulting system may consist of excess oil,
excess water, and a third phase containing a mixture of surfac-
tant molecules, oil, and water organized into liquid crystalline
or bicontinuous microemulsion structures. To form nanoemul-
sions it is important to control the overall system composition
so that only this third phase is formed at the PIT. At tempera- Figure 8 Temperature dependence of the turbidity of tetradecane oil-in-water
nanoemulsions. The turbidity increases above the PIT due to droplet coalescence
tures sufficiently greater than the PIT (≈ T > PIT + 20◦ C), the and phase inversion.
surfactant is more soluble in oil than water, its packing param-
eter is larger than unity, and the formation of a W/O emulsion
is favored. A nanoemulsion can be formed spontaneously by sion was then measured as it was heated at a controlled rate.
rapidly cooling an emulsion from a temperature at or slightly At T << PIT, the turbidity was low because the oil droplets in
above the PIT to a temperature well below the PIT with contin- the nanoemulsion were so small that they did not scatter light
uous stirring. Recently, it has been proposed that the formation strongly. As the temperature approached the PIT, there was an
of nanoemulsions using the PIT method has a similar physico- increase in the turbidity because the optimum curvature tended
chemical basis as their formation using the spontaneous emulsi- towards unity, which increased the droplet coalescence rate and
fication method (Anton et al., 2009). Above the PIT, a non-ionic led to an increase in particle size. Above the PIT, the system was
surfactant/oil/water mixture consists of a W/O emulsion with the highly turbid since a water-in-oil emulsion was formed with rela-
surfactant molecules being present predominantly within the oil tively large droplets that scattered light strongly. If this emulsion
droplets because they are more oil-soluble than water-soluble at was again cooled rapidly to a temperature well below the PIT,
this temperature. When this system is quench-cooled below the a nanoemulsion could be formed again. This figure highlights
PIT, the surfactant molecules rapidly move from the oil phase one of the major limitations of the PIT method of producing
into the aqueous phase (just like the movement of water-miscible nanoemulsions—the droplets tend to be highly prone to droplet
solvent in the spontaneous-emulsification method), which leads coalescence when the temperature of the system is raised. This
to the spontaneous formation of small oil droplets because of could be a problem in many food and beverage applications that
the increase in interfacial area and interfacial turbulent flow require some form of thermal treatment, such as pasteurization,
generated (Fig. 7). sterilization, or cooking. Recently, we developed an approach
An example of the phase behavior of a mixture consisting to overcome this problem by forming nanoemulsions using a
of oil (20 wt% tetradecane), water (74%), and non-ionic sur- non-ionic surfactant with a relatively low PIT (Brij 30), and
factant (6% Brij30) upon heating is shown in Fig. 8 (Rao and then diluting the resulting nanoemulsions in a solution contain-
McClements, 2010). Initially, a nanoemulsion (d = 60 nm) was ing another surfactant (SDS or Tween 80) with none or a high
formed by heating this mixture above the PIT (≈ 36◦ C) and PIT (Rao and McClements, 2010). This second surfactant par-
then rapidly cooling it to 13◦ C. The turbidity of this nanoemul- tially replaced some of the original surfactant from the droplet
FOOD-GRADE NANOEMULSIONS 297

Figure 9 Schematic representation of proposed mechanism for low-intensity emulsification by the catastrophic phase inversion method. The amount of water
added to a W/O emulsion is progressively increased, until a phase inversion occurs and an O/W emulsion is formed.

surfaces and increased the repulsive interactions between the emulsion to another (e.g., W/O to O/W or vice versa) is through
droplets, thereby inhibiting coalescence. a catastrophic-phase inversion, rather than a transitional-phase
Phase Inversion Composition Methods. The phase-inversion inversion as with the PIC or PIT methods (Fernandez et al., 2004;
composition or PIC method is somewhat similar to the PIT Thakur et al., 2008). A transitional-phase inversion occurs when
method, but the optimum curvature of the surfactant is changed the surfactant properties are altered by adjusting a formulation
by altering the formulation of the system, rather than the tem- variable, such as the temperature, pH, or ionic strength. On the
perature (Anton et al., 2009). For example, an O/W emulsion other hand, a catastrophic-phase inversion occurs when the ratio
stabilized by an ionic surfactant can be made to phase invert to of the oil-to-water phases is altered while the surfactant proper-
a W/O emulsion by adding salt. In this case, the packing param- ties remain constant. Practically, catastrophic-phase inversion is
eter is adjusted from p < 1 to p > 1 due to the ability of the usually induced by either increasing (or decreasing) the volume
salt ions to screen the electrical charge on the surfactant head fraction of the dispersed phase in an emulsion above (or below)
groups (Maestro et al., 2008). Alternatively, a W/O emulsion some critical level. For example, a W/O emulsion consisting of
containing a high salt concentration can be converted into an water droplets dispersed in oil is initially formed using a partic-
O/W emulsion by diluting it in water so as to reduce the ionic ular surfactant, and then increasing amounts of water are added
strength below some critical level. Another way to prepare na- to the system with continued stirring (Fig. 9). At low amounts of
noemulsions using the PIC method is by changing the pH to added water, additional water droplets are formed within the oil
alter the electrical charge and stability of emulsions. Fatty acids phase. However, once a critical water content is exceeded, the
may stabilize W/O emulsions at low pH (pH < pKa ) because the coalescence rate of water droplets exceeds the coalescence rate
carboxyl groups are uncharged so they have a relatively high oil of oil droplets, and so phase inversion occurs from a W/O to an
solubility and p > 1. However, they may stabilize O/W emul- O/W system. The value of the critical concentration where phase
sions at high pH values because the carboxyl group becomes inversion occurs, as well as the size of the oil droplets produced,
ionized so they become more water-soluble and p < 1. Conse- depends on process variables, such as the stirring speed, the rate
quently, nanoemulsions can be formed by increasing the pH of a of water addition, and the emulsifier concentration (Thakur et
fatty acid-oil-water mixture from below to above the pKa value al., 2008). The emulsifiers used in catastrophic-phase inversion
of the carboxyl groups (Solè et al., 2006; Maestro et al., 2008). are usually limited to small molecule surfactants that can stabi-
Emulsion Inversion Point (EIP) Methods. In the emulsion lize both W/O emulsions (at least over the short term) and O/W
inversion point (EIP) methods the change from one type of an emulsions (over the long term).
298 D. J. McCLEMENTS AND J. RAO

Figure 10 Formulation-composition map for a typical surfactant-oil-water system.

Rationalizing Phase-Inversion Methods using the HLD Con- oil-water system (Leal-Calderon et al., 2007). It takes into ac-
cept. Recently, a new conceptual approach has been developed count the influence of oil-phase properties (such as oil type),
to rationalize and generalize the phase-inversion behavior ex- aqueous-phase properties (such as salt or alcohol content), and
hibited by surfactant-oil-water (SOW) mixtures (Salager et al., environmental factors (such as temperature) into the relative
2004; Queste et al., 2007). This approach involves describing affinity of a surfactant for the oil and water phases. The HLD
the behavior of a SOW system in terms of a formulation- system is therefore more comprehensive than the traditional
composition map (Fig. 10). The x-axis represents changes in hydrophilic-lipophilic balance (HLB) system that is commonly
the “composition” of the emulsion, which is usually expressed used to describe surfactant behavior, since the HLB system only
as the water-to-oil ratio (WOR = mass of water/mass of oil). usually takes into account the properties of the surfactant itself.
The y-axis represents changes in the “formulation” of the SOW The HLD number of a SOW system can be calculated us-
system, which is usually expressed as the hydrophilic-lipophilic ing the following empirical equation for ethoxylated non-ionic
deviation (HLD). The HLD is a measure of the relative affin- surfactants (Queste et al., 2007):
ity of the surfactant for the hydrophilic (water) phase and the
lipophilic (oil) phase (Salager et al., 2005; Witthayapanyanon HLD = [α − k × EACN – EON]
et al., 2008). Sometimes the surfactant-affinity difference (SAD)
is used to represent the formulation variable rather than the + [a × CA + b × Cs + c × (T − T0 )] (1)
HLD. These two terms are closely related to each other: HLD =
SAD/RT, where R is the gas constant and T is the absolute
temperature. The formulation-composition map is a convenient The first three parameters in this equation are mainly deter-
means of specifying the types of emulsions that can be formed mined by the molecular characteristics of the surfactant and oil
by a given SOW system when either the surfactant properties used, while the last three parameters are mainly determined by
are altered (e.g., by altering pH, ionic strength, or temperature) the influence of system conditions (composition and tempera-
or the overall system composition is changed (e.g., by changing ture) on surfactant properties. Here, α and k are constants that
the relative amounts of oil and water present). depend on the surfactant type; EACN is the equivalent alkane
The HLD value is a dimensionless parameter that character- carbon number of the oil phase, and EON is the ethylene ox-
izes the behavior of a surfactant within a particular surfactant- ide number of the surfactant head group. In addition, CA is the
FOOD-GRADE NANOEMULSIONS 299

concentration of any alcohol present in the system (wt%), a is form reverse micelles or microemulsions in the oil phase;
a constant that depends on the type of alcohol present, CS is (iii) The surfactant tends to stabilize W/O (rather than O/W)
the concentration of any salt present in the system (wt%), b is a emulsions. The more positive the HLD number, the greater
constant that depends on the type of salt present, T is the temper- the affinity for the oil phase.
ature, T0 is a reference temperature (◦ C), and c is a constant that
depends on how the properties of a particular surfactant change Knowledge of the HLD number, the water-to-oil ratio
with temperature. Typically, the reference temperature is 25◦ C. (WOR), and the formulation-composition map for a particu-
Values for the other parameters have been determined for a lar system can be used to rationalize its behavior under different
number of surfactant-oil-water systems. For example, EACN conditions (Fig. 10). The formulation-composition map can be
and other parameters have recently been reported for a number conveniently divided into a number of different regimes, which
of terpenes (Bouton et al., 2009) and for a number of triacyl- are designated by a letter and a sign (Leal-Calderon et al., 2007).
glycerol oils (Phan et al., 2010). In general, the HLD number The sign determines the influence of formulation (HLD num-
of a particular class of non-ionic surfactants tends to increase ber) on the type of emulsion formed. In negative regimes (A-,
with increasing temperature, increasing salt concentration, de- B-, and C-), HLD < 0, and the surfactant favors the formation
creasing EACN, and increasing EON. For certain surfactant-oil of O/W emulsions, whereas in positive regimes (A+, B+, and
combinations the HLD number can be made to change from neg- C+), HLD > 0, and the surfactant favors the formation of W/O
ative to positive as the temperature is increased, which favors emulsions (Salager et al., 2005; Witthayapanyanon et al., 2008).
transition from a O/W to W/O emulsion (see below). The tem- The letter determines the influence of composition (WOR) on
perature at which this transition occurs is the phase-inversion the type of emulsion formed: A refers to a system where the oil
temperature (PIT) mentioned earlier. Triacyglycerol oils (MCT and water phases have fairly similar amounts (WOR ≈ 1) and
and LCT) tend to have much higher EACN numbers than alka- so formation of both O/W and W/O emulsions is equally favor-
nes or flavor oils which accounts for the fact that it is difficult to able; B refers to a system where the oil phase is in excess and
use the PIT method to form nanoemulsions from these oils—the so the formation of W/O emulsions is favored; and C refers to a
temperature required to make the HLD number go from negative system where the water phase is in excess and so the formation
to positive is too high to be reached under normal circumstances. of O/W emulsions is favored. If both the formulation variable
For ionic surfactants the HLD number can be calculated from (HLD number, sign) and the composition variable (WOR, let-
the following expression (Salager at al 2005): ter) favor the formation of a particular emulsion type (e.g., O/W)
then this emulsion is said to be “normal” and will tend to be
HLD = [α  − k  × EACN] + [a  × CA + b × ln(Cs ) stable (Mira et al., 2003; Rondón-Gonzaléz et al., 2006). On the
other hand, if the formulation variable favors one emulsion type
+ c × (T − T0 )] (2) (e.g., W/O) while the composition variable favors the other type
(e.g., O/W), then this system is said to be “abnormal” and will
Once these relationships have been established for a particu- tend to be unstable. In the formulation-composition map shown
lar surfactant-oil-water combination, they can be used to predict in Fig. 10 there are four regimes where the emulsions should be
how the behavior of the surfactant will change when the system stable (A-, A+, B+, and C-) and two regimes where they should
composition (e.g., salt, alcohol, and oil) and environmental con- be unstable (B- and C+). The emulsions formed in the abnormal
ditions (e.g., temperature) are altered. This information can be regimes are usually highly unstable to droplet coalescence and
used to determine the optimum conditions required to form sta- phase separation, but multiple emulsions may be formed near
ble emulsions or nanoemulsions using phase-inversion methods. certain phase boundaries. For example, O/W/O emulsions may
The relationship between the HLD number of a surfactant be formed near the B- to A- boundary, whereas W/O/W emul-
and its ability to stabilize emulsions is highlighted below: sions may be formed near the C+ to A+ boundary (Fig. 10).
The formulation-composition map can be used to describe
both transitional- and catastrophic-phase inversions (Leal-
• HLD < 0: The surfactant has: (i) A higher affinity for the water Calderon et al., 2007). For example, if one moves in a horizontal
phase than the oil phase; (ii) The surfactant tends to form direction from Regime B- to Regime A- by increasing the WOR,
normal micelles or microemulsions in the aqueous phase; the system would undergo a catastrophic-phase inversion from
(iii) The surfactant tends to stabilize O/W (rather than W/O) a W/O to an O/W emulsion, passing through an intermediate
emulsions. The more negative the HLD number, the greater regime where a multiple emulsion (O/W/O) was formed. Con-
the affinity for the water phase. versely, if one moves in a vertical direction from Regime A+ to
• HLD = 0: The surfactant has: (i) An equal affinity for the water Regime A- by decreasing the HLD value from a positive to nega-
and oil phases; (ii) The surfactant tends to form bicontinuous tive value (e.g., decreasing temperature from above to below the
microemulsion or liquid crystalline phases; (iii) The surfactant PIT), the system would undergo a transitional-phase inversion
tends to stabilize neither O/W nor W/O emulsions. from a O/W to a W/O emulsion, passing through an intermediate
• HLD > 0: The surfactant has: (i) A higher affinity for the regime where a bicontinuous system or liquid-crystalline phase
oil phase than the water phase; (ii) The surfactant tends to was formed. This concept is therefore an extremely valuable one
300 D. J. McCLEMENTS AND J. RAO

Figure 11 Various “finishing techniques” can be used to modulate the properties of nanoemulsions after they have been prepared in order to change their
stability or functional properties.

for rationalizing the conditions where it is possible to form stable nanoemulsion for different applications. Interfacial properties
nanoemulsions using either type of phase-transition mechanism. may be altered after nanoemulsion formation using a number of
However, it does not indicate the potential size of the droplets approaches:
formed in each region, i.e., whether a nanoemulsion or an emul-
sion is formed. This will depend on the total amount of the emul- • Interfacial Displacement: A nanoemulsion may be mixed
sifier present, as well as the preparation method used to form the with a solution containing an emulsifier that competes at
emulsion (e.g., the energy intensity for a high-energy method). the oil-water interface thereby displacing some or all of the
original emulsifier from the droplet surfaces (Rao and Mc-
Finishing Techniques Clements, 2010). The interfacial composition will depend on
the relative concentrations and surface activities of the various
After a nanoemulsion has been formed using either a low- emulsifiers present in the system (McClements, 2005).
• Interfacial Deposition: The interfacial characteristics of the
energy or high-energy approach, it is possible to change the
droplet characteristics using a variety of methods, which we droplets in a nanoemulsion may also be changed using inter-
refer to as finishing techniques (Fig. 11). facial deposition methods (Guzey and McClements, 2006a;
McClements, 2010). In this case, the droplets in a nanoemul-
sion are mixed with a solution containing a substance that
Interfacial Engineering
adsorbs to the surface of the existing emulsifier layer. For
For some practical applications of nanoemulsions, it may be example, if the original droplets are coated by an emulsifier
advantageous to alter the nature of the interfacial layer surround- with an electrical charge, then oppositely charged substances
ing the oil droplets after formation. For example, one type of (such as polyelectrolytes or mineral ion) can be deposited onto
emulsifier may be particularly effective at producing very small their surfaces using electrostatic deposition (Guzey and Mc-
droplets during nanoemulsion preparation, but may not be that Clements, 2006a; McClements, 2010). This technique can be
effective at stabilizing the droplets after they have been formed. used to create multilayers around lipid droplets that have dif-
This is true for nanoemulsions produced using the PIT method, ferent functional properties than the original emulsifier layer.
where the non-ionic surfactants used to form the emulsions are • Interfacial Cross-linking: For certain types of emulsifiers it is
not effective at preventing subsequent droplet coalescence at possible to cross-link them after they have adsorbed to the oil-
elevated temperatures (Rao and McClements, 2010). Alterna- water interface, e.g., physically, chemically, or enzymatically.
tively, changing the emulsifier type after the formation of a This approach has been used to covalently cross-link adsorbed
nanoemulsion may be useful in changing interfacial properties protein layers (Dickinson, 1997; Bos and van Vliet, 2001;
such as composition, electrical charge, rheology, or thickness, Sandra et al., 2008) and polysaccharide layers (Littoz and
which may enable one to tune the functional performance of a McClements, 2008).
FOOD-GRADE NANOEMULSIONS 301

Solvent Displacement/Evaporation and a protein emulsifier (whey protein isolate, WPI). The size
of the droplets was then reduced by diluting these emulsions
The size of the droplets in an emulsion or nanoemulsion can
with pure water to promote movement of ethyl acetate into the
be reduced after an initial homogenization step using solvent-
aqueous phase (solvent displacement) and/or by heating them
displacement and/or solvent-evaporation methods (Horn and
to remove the ethyl acetate (solvent evaporation). The droplet
Rieger, 2001; Tan et al., 2005a,b; Chu et al., 2007b). In these
size produced could be controlled using these solvent removal
methods, an oil-in-water emulsion is initially formed by ho-
approaches by varying the ratio of ethyl acetate to corn oil in the
mogenizing an organic phase (lipid and organic solvent) with
initial organic phase: the higher the fraction of ethyl acetate ini-
an aqueous phase (water and hydrophilic emulsifier). The or-
tially present in the organic phase, the smaller the final droplet
ganic solvent is selected based on its water miscibility, boiling
size produced (Fig. 12).
point, safety, and legal status. The lipid may be a liquid oil
One limitation of this approach is that the volume of a droplet
(such as fish, corn, sunflower, or olive oil), a solid fat (such
is proportional to its diameter cubed (V ∝ d 3 ), so that to half the
as hydrogenated palm oil), a lipophilic nutraceutical (such as
droplet diameter one must reduce the droplet volume to 1/8th of
β-carotene, lycopene, or curcumin), or a mixed system, but the
the original value. This means that a large amount of organic sol-
lipid must be capable of being dissolved within the organic
vent must be used compared to the amount of the encapsulated
solvent prior to homogenization. This is because it is difficult
component. The solvent removal method has been successfully
to form emulsions when the oil phase contains any crystalline
used to prepare dispersions of β-carotene nanoparticles from
material. The organic solvent may be hydrophilic, lipophilic,
food-grade ingredients (Tan and Nakajima, 2005a; 2005b; Chu
or amphiphilic (Horn and Rieger, 2001). After homogenization,
et al., 2007a). Various food-grade emulsifiers can be used to
the oil droplets within the emulsion contain a mixture of lipid
prepare nanoemulsions based on this approach, including whey
and organic solvent. In the solvent-displacement method, an am-
proteins, caseinate, and small molecule surfactants (Tan et al.,
phiphilic organic solvent that is partially miscible with water is
2005a,b; Chu et al., 2007a).
used, and the aqueous phase of the emulsion is initially saturated
with the solvent. When this emulsion is diluted with water, then
the organic solvent moves from the oil droplets into the aqueous
Lipid Phase Exchange
phase, which causes the droplets to shrink in size. If necessary,
the remaining organic solvent can then be removed by evap- The composition of the oil phase within the droplets in an
oration. In the solvent-evaporation method, a volatile organic O/W nanoemulsion may be changed after it has been prepared
solvent that is either immiscible (lipophilic solvent) or partially by mixing it with another emulsion or microemulsion system
miscible (amphiphilic solvent) with water can be used (Horn (Fig. 11). When an oil molecule has a finite solubility in the
and Rieger, 2001). In this case, the organic solvent is removed aqueous phase, then it will be transported between lipid droplets
from the oil droplets by evaporation, which causes the droplets through molecular diffusion due to an entropy of mixing effect
to shrink in size. The final droplet size produced using these (Pena and Miller, 2002; Capek, 2004). Alternatively, if the oil
two methods depends on the initial droplet size in the original molecules can be solubilized within surfactant micelles then
emulsion, the amount of organic solvent present, and the water- they may also be transported between the lipid droplets (Weiss
solubility of the organic solvent. Some studies using the solvent et al., 2000). Consequently, it is possible to mix an emulsion
evaporation or displacement method are reported in Table 8. containing droplets with one composition, with another
Our laboratory recently used this approach to form na- emulsion or microemulsion containing droplets of a different
noemulsions containing protein-coated oil droplets (Lee and composition. These compositional ripening effects will then
McClements, 2010). An oil-in-water emulsion was formed us- lead to the transport of oil between droplets until a steady-state
ing a lipid phase that consisted of corn oil and ethyl acetate (an condition is reached (McClements and Dungan, 1993). This
amphiphilic solvent), and an aqueous phase that contained water method could be used to change the oil composition of the

Table 7 Reported applications of nanoemulsion in the food industry found on the internet (2010)

Food product name Company Product description

Canola Active Oil Shemen, Haifa, Israel Nanoencapsulation of fortified phytosterols


Fortified Fruit Juice High vive, USA Nanoencapsulation of fortified vitamin, lycopene, Theanine, and
SunActive Iron
NanoResveratrolTM Life Enhancement, USA plant-based lipid, such as a solid triglyceride (fat), encased by a shell
consisting of a natural phospholipid, such as phosphatidylcholine
delivery system
Nutri-NanoTM CoQ-10 3.1x Softgels Solgar, USA 30 nm O/W system

R 
R
Spray for Life Vitamin Supplements Health Plus International , Inc., USA Nanoencapsulation of fortified vitamin beverage
“Daily boost” Jamba Juice Hawaii, USA Nanoencapsulation of fortified vitamin or bioactive components beverage
“Color emulsion” Wild Flavors, Inc, USA Beta-carotenal, apo-carotenal, or paprika nanoemulsions
302 D. J. McCLEMENTS AND J. RAO

Table 8 An overview of research articles where the properties in O/W nanoemulsions have been altered using finishing techniques, i.e., solvent displacement
of evaporation

Bioactive Particle Diameter Particle Diameter


Oil phase component Surfactant/Co-surfactant Before Finishing After Finishing References

Corn oil, Ethyl acetate — WPI 174 nm < 70 nm Lee and McClements, 2010
Hexane β−carotene Sodium caseinate 1000 nm 18 nm Chu et al., 2007
Acetone β−carotene Sodium caseinate Tween 20, — 30–206 nm (It depends on Yin et al., 2009
decaglycerol monolaurate, sucrose surfactant type)
fatty acid ester
Acetone β−carotene Gelatin or Tween 20/polylactic acid — 74 nm (gelatin) Ribeiro et al., 2008
77 nm (Tween 20)
Acetone β−carotene Tween 20 — 9–280 nm Silva et al., 2010
Hexane Hexane Tween 20 — 60–140 nm Tan and Nakajima, 2005

droplets in a nanoemulsion, e.g., an oil-soluble antioxidant, than when it is in its bulk state because of supercooling and cur-
antimicrobial, or nutritional component could be introduced vature effects (Turnbull and Cormia, 1961; Westesen, 2000).
into a nanoemulsion after it has been formed.

Lipid Crystallization DESIGNING FUNCTIONAL NANOEMULSION


PARTICLES
The oil phase used in the preparation of nanoemulsions by ei-
ther the high-energy or low-energy methods must usually be liq- The characteristics of the particles in nanoemulsions
uid. Nevertheless, it is possible to change the physical state of the ultimately determine the bulk physicochemical and functional
oil phase after a nanoemulsion has been formed (Fig. 11), e.g., properties of nanoemulsion-based products, e.g., their optical
by crystallizing the lipid phase (Weiss et al., 2007). The oil phase properties, stability, rheology, digestibility, and release charac-
can be fully or partially crystallized after nanoemulsion forma- teristics. In this section, we highlight the most important particle
tion by controlling its composition and the temperature-time his- characteristics that can be controlled to obtain nanoemulsions
tory of the product. In this case, a lipid phase is used that is crys- with specific functional properties.
talline at the final product temperature. Before a nanoemulsion
can be produced the lipid phase has to be heated above its melt-
ing point and maintained at this temperature throughout the man- Particle Composition
ufacturing process to ensure that it remains liquid. The resulting
nanoemulsion is then cooled at a controlled rate to crystallize the The composition of the particles in nanoemulsions can be
oil droplets, without promoting product instability. By careful controlled by appropriate selection of ingredients and process-
selection of oil type and thermal history, it is possible to control ing operations. The particles in oil-in-water nanoemulsions can
the type, concentration, and location of the crystals within the conveniently be considered to consist of a core of lipophilic
droplets, which can lead to novel functional properties (Bum- material surrounded by a shell of surface-active material (Mc-
mer, 2004; Kesisoglou et al., 2007; Weiss et al., 2007). It should Clements, 2005). A number of different non-polar food ingredi-
be noted that fat crystallization usually occurs at a much lower ents can be used to create the lipophilic core, including triacyl-
temperature when an oil is divided into nanometer-sized droplets glycerols, diacylgycerols, monoacylglycerols, free fatty acids,

Figure 12 Nanoemulsions can be prepared by the homogenization/solvent evaporation method. An emulsion is prepared with an oil phase that contains a
mixture of oil and solvent. The solvent is then removed by evaporation, which causes the droplets to shrink in size. The photographs show whey-protein stabilized
5% oil-in-water emulsions before and after solvent evaporation: oil phase = 5% corn oil: 95% ethyl acetate.
FOOD-GRADE NANOEMULSIONS 303

Figure 13a Predicted influence of droplet size and shell layer thickness on the effective composition of core-shell particles in nanoemulsions. As the core size
decreases or the shell layer thickness increases, the fraction of core in the particles decreases.

flavor oils, essential oils, mineral oils, fat substitutes, waxes, r is the radius of the core, and δ is the thickness of the shell
weighting agents, oil-soluble vitamins, and nutraceuticals (such (Fig. 1). The dependence of
S on r for nanoemulsions with
as carotenoids, curcumin, phytosterols, and Co-enzyme Q). As different core radii and shell-layer thicknesses is shown in Fig.
mentioned earlier, these ingredients have different densities, 13a. This figure shows that the composition of the particles in
refractive indices, polarities, viscosities, and melting behavior, small nanoemulsions may be quite different from those in con-
which will impact the functional performance of the overall sys- ventional emulsions made from the same ingredients due to the
tem (Table 1). A variety of surface-active food ingredients can be large contribution of the shell. For example, a particle with a
used to create the shell surrounding the lipophilic core, including core radius of 25 nm and a shell thickness of 10 nm is about 37%
surfactants, phospholipids, proteins, polysaccharides, minerals, core and 63% shell. The core-shell model used to calculate these
and solid particles. The choice of these components plays a curves certainly over-simplifies the actual structure of the parti-
critical role in determining particle properties, such as physical cles found in real nanoemulsions. In reality, the shell is not ho-
stability, interactions, release characteristics, and digestibility. mogeneous but consists of a complex mixture of different kinds
In conventional emulsions, the shell thickness (δ) is much of molecules that accumulate at the oil-water interface, such as
smaller than the core radius (r), i.e., δ << r, and so the core- proteins, phospholipids, small molecule surfactants, bioactive
shell particles consist primarily of the oil phase. On the other components, mineral ions, water, lipids, and polymers (Fig. 1).
hand, in nanoemulsions, the shell thickness may be of the same In addition, its structure may be highly dynamic as molecules
order of magnitude as the core radius (δ ≈ r), and hence the shell move in and out of the interface, and undergo positional and
layer makes an appreciable contribution to the overall particle conformational changes at the interface. The precise structure
composition (Tadros et al., 2004; Mason et al., 2006). This effect and orientation of these molecules at the particle surface may
can be described by the following equation: be important in many practical applications.
The size-dependence of particle composition has a number
r3 of important practical implications. First, the dependence of

S = 1 − (3) particle composition on particle size will impact those particle
(r + δ)3
properties (such as refractive index, density, and permeability)
that alter the bulk physicochemical properties and stabilities
Here
S (= Vshell /Vtotal ) is the volume fraction of the shell of nanoemulsions (such as optical properties, creaming rate,
compared to the total particle, Vshell is the volume of the shell, and release characteristics). Second, the loading capacity of a
Vtotal is the total volume of the entire particle ( = Vcore + Vshell ), lipophilic component within the particles in a nanoemulsion may
304 D. J. McCLEMENTS AND J. RAO

decrease as the core size decreases because of the reduction in For sterically-stabilized systems, the thickness of the shell
the fraction of lipophilic core present (Fig. 13a). Third, the de- is determined primarily by the dimensions of the layer of ad-
pendence of particle composition on particle size may influence sorbed surface-active molecules. For electrostatically-stabilized
the accuracy of data obtained from instrumental techniques de- systems, the effective shell thickness (δ) may be considerably
signed to measure particle size, such as laser diffraction, spectro- greater than the physical dimensions of the adsorbed layer
turbidity, and dynamic light scattering (McClements, 2005). The (Tadros et al., 2004; Mason et al., 2006). In these systems,
mathematical models used by these instruments to calculate the the thickness depends on the range of the electrostatic interac-
particle size distribution from some measured physical prop- tions (Debye screening length, κ −1 ), which depends on the ionic
erty (such as a light-scattering pattern) usually assume that the strength of the surrounding aqueous phase (McClements, 2005). √
particles are homogeneous spheres with well-defined properties For aqueous solutions at room temperatures, κ −1 ≈ 0.304/ I
(such as refractive index and density). Consequently, there may nm, where I is the ionic strength expressed in moles per liter
be some errors in the reported particle size distributions in na- (Israelachvili, 1992). For example, the ranges of the electro-
noemulsions where the particles are core-shell structures rather static interactions in monovalent salt solutions with different
than simple homogeneous spheres. ionic strengths are: κ −1 = 0.3, 0.96, 3, 9.6, and 30 nm for I =
1, 10−1 , 10−2 , 10−3 ,and 10−4 M, respectively. At low ionic
strengths, the Debye screening length may therefore be on the
Particle Concentration
same order as the particle radius in nanoemulsions, which can
have a major impact on the bulk physicochemical properties of
In general, the particle concentration in a colloidal dispersion
the system, such as the optical, rheological, and stability prop-
is usually expressed as the number, mass, or volume of parti-
erties (Weiss et al., 2000; Weiss and McClements, 2001; Mason
cles per unit volume or mass of the total system (McClements,
et al., 2006; Kawada et al., 2010). The impact of this effect
2005). In conventional oil-in-water emulsions, the particle con-
on the rheology of nanoemulsions is discussed in section titled
centration is usually characterized in terms of the disperse phase
“Rheological Properties”.
volume fraction (φ), which is the volume of oil per unit volume
The particle concentration within a nanoemulsion is usually
of emulsion. In oil-in-water nanoemulsions, the effective par-
controlled during the manufacturing process by controlling the
ticle concentration may be quite different from the fraction of
proportions of the two immiscible liquids (oil and water) used to
pure disperse phase present, which can have a major impact on
prepare it. However, the particle concentration can also be varied
their physicochemical and functional attributes (Tadros et al.,
after a nanoemulsion formation using dilution (e.g., by adding
2004; Mason et al., 2006).
more continuous phase) or concentration (e.g., by gravitational
As discussed in the previous section, the particles in
separation, filtration, centrifugation, or evaporation) methods. It
emulsions can be considered to have a core-shell structure (Fig.
should be noted that concentrating nanoemulsions by many of
1). Hence, the overall effective volume fraction (φ effective ) is the
these methods is much more difficult than concentrating con-
sum of the volume fractions of the core (φ core ) and the shell
ventional emulsions because of their very small droplet size.
(φ shell ): φ effective = φ core + φ shell . An expression for the effective
As discussed previously, the effective size of the particles in
volume fraction can be obtained from knowledge of the particle
a nanoemulsion may also be changed by using emulsifiers that
dimensions:
form different adsorbed layer thicknesses, or by altering solution
  conditions that impact the range and magnitude of electrostatic
δ 3
φeffective = φcore 1 + (4) interactions between charged particles (such as the pH or ionic
r strength).

In conventional emulsions, the shell thickness is much smaller


Particle Dimensions
than the oil core (δ << r), and so the effective droplet concen-
tration is similar to the oil-core concentration (φeffective ≈ φcore ). The dimensions of the particles within a nanoemulsion are
However, in nanoemulsions the thickness of the emulsifier layer one of its most important characteristics since they influence
may approach that of the droplet radius (δ ≈ r) and so the effec- its optical, rheological, stability, and release characteristics, as
tive particle volume fraction may be appreciably higher than the well as potentially altering its biological fate. The dimensions of
oil phase volume fraction (φeffective >> φcore ). The influence of the particles within a colloidal dispersion are usually character-
the shell layer on the effective particle volume fraction is shown ized by the particle-size distribution (PSD), which represents the
in Fig. 13b for nanoemulsions with different oil core sizes and fraction of particles within a range of discrete size classes (Wal-
shell layer thicknesses. These calculations indicate that the stra, 2003; McClements, 2005). A PSD is typically presented as
effective volume fraction of the particles in nanoemulsions may a table or plot of particle concentration (e.g., volume or number
be many times higher than the actual volume fraction of the percent) versus particle size (e.g., diameter or radius). For many
oil phase. This increase in effective particle concentration may purposes, it is more convenient to represent the full PSD using
cause appreciable changes in the bulk physicochemical prop- a measure of the central tendency of the distribution (e.g., mean
erties of nanoemulsions, such as rheology and stability. (See or median) and a measure of the width of the distribution (e.g.,
section titled “Physicochemical Properties of Nanoemulsions.”) standard deviation or polydispersity index).
FOOD-GRADE NANOEMULSIONS 305

Figure 13b Predicted influence of droplet size and shell layer thickness on the effective concentration of core-shell particles in nanoemulsions. As the core size
decreases or the shell layer thickness increases, the effective particle concentration increases.

The effective radius of the particles in an oil-in-water na- The PSD of an emulsion can usually be controlled by vary-
noemulsion is the sum of the radius of the lipophilic core and ing nanoemulsion preparation conditions and/or system com-
the thickness of the shell: reffective = r + δ. There are appre- position. For example, in high-energy methods the droplet size
ciable differences in the thicknesses of the layers formed by depends on the intensity and duration of the energy input, the
food-grade emulsifiers—typically, small molecule surfactants type and concentration of the emulsifier used, the interfacial
(such as Tweens and Spans) < globular proteins (such as egg, tension, and the relative viscosities of the disperse and contin-
whey, or soy proteins) < flexible proteins (such as caseinate or uous phases (McClements, 2005; Jafari et al., 2006). Smaller
gelatin) < polysaccharides (gum Arabic or modified starch). In droplets can usually be produced by increasing the intensity or
addition, the thickness of the shell layer may be increased by duration of homogenization, by increasing the concentration of
adsorbing additional material to the droplet surfaces after they the emulsifier used, or by controlling the viscosity ratio (Wal-
have been formed, e.g., by electrostatic deposition of one or stra, 1993; Schubert and Engel, 2004; Wooster et al., 2008). In
more layers of charged biopolymers (or other charged species) low-energy methods, the droplet size depends on factors such as
onto the surfaces of oppositely charged emulsifier-coated oil system composition (such as surfactant-oil-water ratio, surfac-
droplets (Guzey and McClements 2006a; 2006b). In reality, it tant type, and ionic strength) and environmental conditions (such
may be difficult to accurately define the physical dimensions as temperature-time history, and stirring speeds) (Anton et al.,
of the particles in a nanoemulsion, since the shell may con- 2008; 2009). The most effective approach to control the size of
sist of a polymeric or electrically-charged layer with no distinct the particles within a specific nanoemulsion therefore depends
edge. In addition, the most appropriate shell thickness to use on the particular homogenization method used to prepare it. Fi-
in any mathematical calculations may depend on the partic- nally, it should be stated that it is often important to stabilize the
ular situation encountered. For example, for a nanoemulsion particles within a nanoemulsion against aggregation or growth
containing charged particles, δ may be taken to be the Debye after they have been formed, in order to prevent undesirable
screening length for rheological purposes, but it may be taken to changes in particle dimensions during storage or utilization.
be the physical thickness of the surfactant head-groups for light
scattering purposes. As discussed above, the fact that the shell Particle Charge
layer makes a significant contribution to the overall particle di-
mensions in nanoemulsions has important consequences for the The droplets in nanoemulsions often have an electrical charge
measurement and interpretation of particle-size distributions. because of adsorption of ionized emulsifiers, mineral ions, or
306 D. J. McCLEMENTS AND J. RAO

biopolymers to their surfaces (McClements, 2005). The sign


and magnitude of the electrical charge on the droplets plays
an important role in the functional performance and stability
of nanoemulsions, e.g., aggregation stability, interaction with
other food components, and ability to adhere to non-biological
(such as packaging materials) and biological surfaces (such as
the tongue, mucous layer, or other locations within the gastroin-
testinal tract). The electrical properties of a droplet are usually
characterized in terms of its surface-charge density (σ ), sur-
face potential (0 ), and/or ζ −potential (ζ ) (Hunter, 1986). The
surface-charge density is the number of unit electrical charges
per unit surface area. The surface-electrical potential is the free
energy required to increase the surface-charge density from zero
to σ by bringing charges from an infinite distance to the sur-
face through the surrounding medium. The surface-electrical
potential depends on the ionic composition of the surrounding
medium due to electrostatic screening effects and usually de-
Figure 14a The electrical charge on oil droplets can be varied by mixing
creases as the ionic strength of the aqueous phase increases. different kinds of ionic and non-ionic surfactants together. In this system we
The zeta-potential (ζ ) is the electrical potential at the “shear made a nanoemulsion containing lipid droplets stabilized by a non-ionic surfac-
plane,” which is defined as the distance away from the droplet tant (Tween 80), and then added different amounts of either an anionic (SDS)
surface below which the counter-ions remain strongly attached or cationic (LAE) surfactant after homogenization to vary droplet charge (data
to the droplet when it moves in an electrical field. Practically, provided by Khalid Ziani). The SDS concentration is expressed as a negative
number for convenience.
the ζ -potential is often a better representation of the electri-
cal characteristics of a droplet than the surface potential be-
acids) (Kralova et al., 2009), whereas those stabilized by cationic
cause it inherently accounts for the adsorption of any charged
surfactants have a positive charge (e.g., lauric arginate) (Asker
counter-ions. In addition, the ζ -potential is much easier to mea-
et al., 2009). It is possible to vary the magnitude of the droplet
sure than the surface potential or charge density, and therefore
charge in surfactant-stabilized systems by using different ratios
droplet charges are usually characterized in terms of ζ - poten-
of ionic and non-ionic surfactants to prepare them (Fig. 14a).
tial (Hunter, 1986). The range of the electrostatic interactions
The droplet charge can also be varied by using biopolymers
in a system is described by the Debye screening length (κ −1 ),
which decreases as the ionic strength of the aqueous solution
surrounding the droplets increases. (See section titled “Particle
Concentration.”)
Analytical instruments based on particle electrophoresis and
electro-acoustic are widely used to measure the ζ -potential of
the droplets in emulsions (McClements, 2005). These instru-
ments utilize mathematical models to convert the measured
electro-dynamic properties of a particle into a ζ - potential. These
models typically make assumptions about the size of the parti-
cles (r) relative to the thickness of the Debye screening length
(κ −1 ), e.g., it is often assumed that κ −1 << r. This assumption
may be valid for the relatively large droplets in conventional
emulsions, but may not be valid for the very small droplets in
nanoemulsions, particularly at low ionic strengths (high κ −1 ). It
is therefore important to take this effect into consideration when
determining the ζ -potential of droplets in nanoemulsions.
The electrical characteristics of nanoemulsion droplets can
be controlled by careful selection of particular emulsifier types.
In principle, droplets stabilized by non-ionic surfactants (e.g.,
Tweens and Spans) should have no droplet charge, but in prac- Figure 14b The electrical charge on oil droplets in nanoemulsions can be
tice they often have a significant negative charge which may be varied by using different kinds of biopolymer emulsifiers or biopolymer coat-
due to the presence of free fatty acids or other ionic impurities ings. This figure shows the electrical charge versus pH profiles for various kinds
remaining from their production or generated during storage of biopolymers in aqueous solutions. Proteins are amphoteric molecules whose
(Mun et al., 2005). Droplets stabilized by anionic surfactants charge goes from positive at low pH to negative at high pH, with a point of zero
charge depending on protein type. Polysaccharides may be anionic or cationic
have a negative charge (e.g., lecithin, DATEM, CITREM, fatty depending on their charge groups.
FOOD-GRADE NANOEMULSIONS 307

with different electrical characteristics (Fig. 14b). Droplets sta- present. Some types of surface-active agents form relatively
bilized by commercial polysaccharide emulsifiers tend to have low-viscosity interfacial layers (e.g., small molecule surfac-
a net negative charge (e.g., gum Arabic and modified starch) tants, flexible biopolymers), whereas others form more viscous
due to the presence of anionic groups (e.g., sulfate or carboxyl) or gel-like layers (e.g., surface- or thermally-denatured globular
on the polymer chains (Chanamai and McClements, 2002). Oil proteins, or cross-linked biopolymers). Interfacial rheology may
droplets stabilized by proteins (e.g., whey protein, casein, soy be important in determining the particle stability to aggregation,
proteins, egg proteins) have a charge that depends on the solu- or in determining the bulk rheology in systems where there is
tion pH relative to the isoelectric point (pI) of the protein (Gu extensive interfacial layer overlap.
et al., 2005). Protein-coated droplets have a net positive charge The interactions of the particles in a nanoemulsion with other
for pH < pI, no net charge at pH = pI, and a negative charge at particles or surfaces can be controlled in a number of ways. Non-
pH > pI. However, it should be stressed that protein molecules specific electrostatic interactions can be controlled by control-
have regions of both positive charge and negative charge on their ling the electrical charge on the particle surfaces as described
surfaces, and this heterogeneous charge distribution can greatly in section titled “Particle Charge”. For example, a positively-
impact their behavior (Cooper et al., 2005). Proteins with differ- charged particle will tend to stick to a negatively-charged sur-
ent isoelectric points can be selected for particular applications, face, and vice versa. The surfaces of nanoparticles can also be
e.g., β-lactoglobulin has a pI near 5, but lactoferrin has a pI near modified to alter their “stealth” character within the human body
8. The charge on emulsifier-coated droplets in nanoemulsions by attaching hydrophilic polymers to them, e.g., polyethylene
can also be altered by adsorption of other charged substances glycol (PEG) (Howard et al., 2008). Particles with highly hy-
onto their surfaces, such as proteins, polysaccharides, phos- drophilic surfaces tend to be less prone to removal by the natural
pholipids, or multivalent ions (McClements, 2005; Guzey and defenses of the bodies, which increases the residence time of the
McClements, 2006a; Johnston et al., 2006; Quinn et al., 2008; particles in the systemic circulation. Particle surfaces may also
Ochs et al., 2010). be made more hydrophobic by using surface-active biopoly-
mers that have some non-polar character, such as surface- or
thermally-denatured globular proteins. Nanoparticles can also
Interfacial Characteristics be engineered so their surfaces have specific ligands attached
that are capable of binding to specific biological entities or or-
Controlling the interfacial characteristics of particles is one of gans within the human body (Hashida et al., 2005; des Rieux
the most powerful methods of designing nanoemulsion systems et al., 2006). These surface-modified nanoparticles can be used
with specific functional attributes (McClements, 2005; Mason for targeted- or sustained-release applications.
et al., 2006). We have already highlighted how some interfa- The environmental responsiveness of the particles in na-
cial characteristics of the particles in nanoemulsions can be noemulsions is also largely determined by their interfacial char-
controlled, e.g., their composition (See section titled “Particled acteristics. Particles may be designed so that they respond in
Composition”), thickness (see section titled “Particle Dimen- specific ways to changes in environmental conditions, such as
sions”), and electrical charge (see the previous section). There temperature, pH, ionic strength, and enzyme activity (Guzey
are also a number of other interfacial properties that may also and McClements, 2006a). For example, a nanoemulsion may be
be important for determining the overall functional attributes of designed so that the entire particle remains intact under a certain
nanoemulsions, such as interfacial permeability, rheology, inter- set of environmental conditions, but that either the shell and/or
actions, and environmental responsiveness. The selective per- core break down under another set of conditions. This feature is
meability of an interfacial layer to different kinds of molecules particularly important in the development of food-grade deliv-
depends on its porosity and specific interactions. Some surface- ery systems based on nanoemulsions where the delivery system
active molecules may form interfacial layers that are so close- must remain stable in the food product, but becomes unstable in
packed and/or rigid that they are able to inhibit the diffusion of a specific region of the GI tract (McClements et al., 2009).
other molecules through them, e.g., enzymes, lipids, and min- The interfacial characteristics of the particles in a nanoemul-
eral ions. On the other hand, other surface-active molecules sion can be controlled by selection of a particular emulsifier
may be able to form interfacial layers that have open and/or type, such as a particular surfactant, phospholipid, protein, or
dynamic structures that enable molecules to easily pass through polysaccharide. For example, small molecule surfactants tend to
them. In principle, it is possible to design interfacial layers form fairly thin fluid-like layers; flexible biopolymers (such as
that selectively allow only certain kinds of molecule through, caseinate) tend to form relatively thick fluid-like layers; whereas
e.g., based on size, morphology, or charge. Consideration of globular proteins (such as whey or soy protein) tend to form
interfacial permeability may be important when designing sys- thin elastic-like layers due to interfacial cross-linking (Dick-
tems that control the release of functional agents from inside inson, 1992; Dickinson, 2003). Interfacial characteristics may
the particles or where it is important to keep two or more also be controlled by using blends of different kinds of emulsi-
chemically-reactive substances isolated from each other. The fiers in combination (Fig. 11), with the blending being carried
rheology of interfacial layers can often be manipulated by con- out either before or after homogenization. Finally, the interfacial
trolling the type and interactions of surface-active molecules characteristics may be altered by depositing successive layers of
308 D. J. McCLEMENTS AND J. RAO

charged biopolymers (or other charged species) onto oppositely- products (McClements, 2002; Mason et al., 2006). Some food
charged lipid droplets to form nanolaminated coatings around products should be clear or only slightly turbid (e.g., clear forti-
them with different thicknesses, charges, or environmental re- fied waters, soft drinks, juices, jellies, jams, deserts, dressings,
sponsiveness (Guzey and McClements, 2006a; Johnston et al., and sauces) and so the addition of any delivery system should
2006; Quinn et al., 2008; Ochs et al., 2010). not make them look cloudy or opaque. Other food products
are optically opaque (e.g., turbid or opaque dressings, mayon-
naise, sauces, dips, creams, and yogurts) and light scattering by
Particle Physical State the oil droplets in a delivery system may make an important
contribution to their overall appearance. In general, the optical
Nanoemulsions are normally prepared from oils that are liq- properties of a colloidal dispersion are characterized in terms
uid, but it is possible to form them from oil phases that are of its opacity and its color, which can be quantified in terms
able to fully or partially crystallize at the final temperature of of tristimulus color coordinates, such as the L ∗ a ∗ b∗ system
application (Müller et al., 2002; Weiss et al., 2008). In these (McClements, 2002; McClements, 2005). In this color system,
cases, the oil phase is kept in liquid state during the forma- L∗ represents lightness, and a∗ and b∗ are color coordinates:
tion of the nanoemulsion, i.e., by ensuring that the temperature where +a∗ is the red direction, −a∗ is the green direction; +b∗
is above the melting point of the lipid. The oil droplets can is the yellow direction, −b∗ is the blue direction; low L∗ is
then be crystallized after their formation by cooling the na- dark and high L∗ is light. The opacity of an emulsion can then
noemulsion sufficiently below the melting temperature of the by characterized by its lightness (L*), and the color intensity
lipid (Walstra, 2003; Muller and Keck, 2004; Wissing et al., of an emulsion can then be characterized by its chroma: C =
2004; McClements, 2005). The above approach has been used (a ∗2 +b∗2 )1/2 . The higher the value of the chroma, the more
to form solid lipid nanoparticles (SLN) or nanostructured lipid intense is the color. The optical properties of conventional emul-
carriers (NLC) (Fig. 11), which are nanoemulsions where the oil sions (which are usually highly turbid or opaque) are typically
phase has been fully or partly solidified (Müller et al., 2002). An characterized using reflectance measurements, while the opacity
SLN contains lipid droplets that are fully crystallized and have is expressed in terms of the lightness (L*). On the other hand, the
a highly-ordered crystalline structure. An NLC contains lipid optical properties of nanoemulsions (which are usually clear or
droplets with a less-ordered crystalline structure or an amor- slightly turbid) are typically characterized using transmission
phous solid structure, which can aid in the encapsulation and measurements, and their opacity is expressed in terms of the
retention of highly lipophilic active components. The crystal- turbidity (τ ).
lization temperature of an emulsified fat in a nanoemulsion may In general, the overall optical properties of nanoemulsions
be appreciably below that of the same fat in a bulk phase because are mainly determined by the relative refractive index, parti-
of supercooling effects (Gibout et al., 2007). In addition, the na- cle concentration, and particle-size distribution (McClements,
ture of the crystals formed in a nanoemulsion may be different 2002; McClements, 2002; McClements, 2005). The turbidity
from those formed by a bulk fat because of curvature affects, or lightness of nanoemulsions tends to increase with increasing
the limited volume present for crystal growth in droplets, and refractive index contrast, with increasing particle concentration,
the lack of secondary nucleation sites (Muller and Keck, 2004; and have a maximum value at a particle size where light scat-
Wissing et al., 2004). The concentration, nature, and location of tering is strongest (i.e., r ≈ λ). The color intensity (chroma) is
the fat crystals within the lipid droplets in an emulsion can be inversely related to the degree of light scattering, i.e., when the
controlled by careful selection of oil type (e.g., solid fat content lightness or turbidity increases, the color intensity decreases. It
versus temperature profile, polymorphic forms), thermal history is for this reason that the color intensity of an emulsion may
(e.g., temperature versus time), emulsifier type (e.g., tail group decrease when the droplet concentration is increased or droplet
characteristics), and droplet size (Muller et al., 2000; Walstra, size is changed (Chantrapornchai et al., 1998; Chantrapornchai
2003; Muller and Keck, 2004). A major potential advantage et al., 1999; Chantrapornchai et al., 2001).
of SLN and NLC systems is that the rate of molecular diffu- The dependence of the specific turbidity on particle diam-
sion through the lipid phase can be reduced, which can slow eter for emulsions containing monodisperse (spread = 1) or
down chemical-degradation reactions and improve the stability polydisperse (spread = 1.5, 2.0, or 2.5) droplets was calculated
of encapsulated lipophilic components (Weiss et al., 2007). using Mie theory at 600 nm (Fig. 15). The oil phase was as-
sumed to have a refractive index of 1.45 and the water phase
of 1.33. The specific turbidity was defined as the predicted tur-
PHYSICOCHEMICAL PROPERTIES bidity divided by the oil droplet concentration (expressed as a
OF NANOEMULSIONS volume percentage). From visual observation, emulsions can
be considered to be optically transparent when the turbidity at
Optical Properties 600 nm is less than about 0.05 cm−1 for a 0.1 wt% oil-in-
water nanoemulsion, which corresponds to a specific turbidity
The impact of droplet characteristics on overall appearance of about 0.5 cm−1 %vol−1 . For monodisperse emulsions, the pre-
is often an important factor to consider when designing food dicted specific turbidity was relatively low (< 0.5 cm−1 %vol−1 )
FOOD-GRADE NANOEMULSIONS 309

for droplets with diameters less than about 80 nm, but increased bution. These calculations also demonstrate that it is important
appreciably when the droplets were larger than this size. As the to prevent the droplets from growing during storage otherwise
particle-size distribution became more polydisperse (increasing the emulsions will become cloudy, e.g., by flocculation, coales-
spread from 1.0 to 2.5), the turbidity versus the particle diam- cence, or Ostwald ripening.
eter relationship broadened. For polydisperse systems, a mean Experimental measurements of the influence of droplet size
particle diameter of < 30 nm (for a spread of 2.5), < 40 nm (for on the turbidity of protein-stabilized nanoemulsions are shown
a spread of 2.0) or < 60 nm (for a spread of 1.5) was required in Fig. 16. The turbidity increment (i.e., the increase in turbid-
for the systems to appear transparent, i.e., for specific turbidity ity per cm per 1% increase in droplets) of the nanoemulsions
< 0.5 cm−1 %vol−1 . This effect can be attributed to the presence increases as the droplet diameter increases, particularly above
of a population of larger particles in broad particle distribu- about 130 nm. For conventional emulsions, the lightness in-
tions that scatter light much more intensely than the smaller creases steeply as the oil droplet concentration is increased from
particles. The precise particle-size dependence of the specific 0 to 5 wt%, but then increases more gradually at higher droplet
turbidity depends on the particle size distribution of the emul- concentrations (Chantrapornchai et al., 1999). For nanoemul-
sion in question. This highlights the importance of controlling sions, the turbidity increases approximately linearly with in-
both the mean particle size and the width of the particle-size dis- creasing droplet concentration for dilute systems, with the slope
tribution in creating products that appear optically transparent. depending on the droplet size (Lee and McClements, 2010).
These theoretical predictions are in agreement with the experi- The theoretical predictions shown in Fig. 15 were based on
mental data of Wooster and co-workers, who reported that the the assumption that the particles in the nanoemulsions were ho-
majority of droplets in a nanoemulsion should have a diameter mogeneous spherical oil droplets with a well-defined refractive
< 80 nm to obtain optically transparent systems (Wooster et al., index. In practice, the particles in nanoemulsions consist of an
2008). In summary, in order to prepare nanoemulsions that are oil core surrounded by a shell of adsorbed material (e.g., emulsi-
suitable for application in transparent food or beverage products fiers) that usually has different optical properties (refractive in-
it is necessary to ensure that the majority of droplets fall below dex) than the oil phase. For the small particles in nanoemulsions
some critical value (d ≈ 80 nm), which can be achieved by this emulsifier layer makes up a relatively large volume fraction
having a small mean diameter and a narrow spread of the distri- of the coated particles (Fig. 1). To a first approximation, the

Figure 15 Calculated dependence of the specific turbidity of oil-in-water nanoemulsions and emulsions on mean particle size for systems with various particle
size distributions. It is assumed that the systems have a log-normal particle size distribution, which is characterized by its geometric mean and spread.
310 D. J. McCLEMENTS AND J. RAO

Figure 16 Turbidity increment versus mean droplet diameter for corn oil-in-water nanoemulsions stabilized by whey protein isolate produced using the
homogenization/solvent evaporation method. Adapted from Lee and McClements (2010).

refractive index of a coated droplet can be given by the weighted 2005; Genovese et al., 2007). Nanoemulsions may have consid-
average of the refractive indices of the different components: erably different rheological properties than conventional emul-
sions with the same oil content due to the relatively small size
nparticle =
S nS + (1 −
S )nC (5) of the droplets. A number of reviews of nanoemulsion rheology
have been published recently, which the reader is referred to for
Where n is the refractive index,
is the volume fraction, and additional information (Tadros et al., 2004; Mason et al., 2006).
the subscripts S and C stand for the shell (emulsifier layer) and The rheology of dilute colloidal dispersions is typically char-
core (oil phase), respectively. The emulsifier layer is usually acterized in terms of their shear viscosity (McClements, 2005;
comprised of surfactant and/or polymer molecules that have a Genovese et al., 2007). For dilute colloidal dispersions (φ <
larger refractive index than either oil or water. Hence, the refrac- 0.05), the viscosity can be described by Einstein’s equation:
tive index of the particles will tend to increase as the emulsifier η
concentration in the interfacial layer increases, as the thickness = 1 + 2.5φ (6)
η0
of the interfacial layer increases, or as the oil droplets become
smaller, which may appreciably alter the overall scattering prop- Here, η is the viscosity of the overall system, η0 is the viscosity
erties of the particles. In practice, a more sophisticated mathe- of the continuous phase, and φ is the disperse phase volume
matical model of light scattering should be used that takes into fraction. This equation is not applicable at higher droplet con-
account the core-shell structure of the particles rather assuming centrations due to droplet-droplet interactions. For concentrated
that they are homogeneous spheres (Quirantes et al., 1997). conventional emulsions, the viscosity can be described by the
following semi-empirical equation, which takes into account the
increase in viscosity that occurs in concentrated systems due to
Rheological Properties droplet-droplet interactions:
 
Foods that contain lipid droplets may exhibit a wide variety η φ −2
of different rheological characteristics (e.g., viscous liquids, vis- = 1− (7)
η0 φc
coelastic liquids, viscoelastic solids, plastics, or elastic solids)
depending on their composition and microstructure (Quemada Here, φ is the disperse phase volume fraction, and φc is a critical
and Berli, 2002; Walstra, 2003; Berli et al., 2005; McClements, disperse-phase volume fraction above which the droplets are so
FOOD-GRADE NANOEMULSIONS 311

closely packed together that they cannot easily flow past each Gravitational Separation
other. Typically, φc has a value of around 0.4 to 0.6 depending on
Gravitational separation is one of the most common forms of
the nature of the system. For conventional O/W emulsions φ is
instability in conventional emulsions, and may take the form of
usually taken to be the volume fraction of the oil phase present,
either creaming or sedimentation depending on the relative den-
but for nanoemulsions φ should be replaced by the effective
sities of the dispersed and continuous phases.Creaming is the
disperse phase volume fraction (φeffective ) given in Eq. (2). As
upward movement of droplets when they have a lower density
discussed earlier, the effective disperse-phase volume fraction
than the surrounding liquid, whereas sedimentation is the down-
of an O/W nanoemulsion may be much greater than the volume
wards movement of droplets when they have a higher density
fraction of the oil phase because of the presence of the adsorbed
than the surrounding liquid. Liquid edible oils (such as triacyl-
layer (Fig. 13b).
glycerol or flavor oils) normally have lower densities than liquid
The viscosity of a conventional emulsion increases with
water and so creaming is more prevalent in conventional oil-in-
increasing droplet concentration, gradually initially and then
water emulsions, whereas sedimentation is more prevalent in
steeply as the droplets become more closely packed (Fig. 17).
water-in-oil emulsions. However, if an oil-in-water emulsion
Around and above the droplet concentration where close pack-
contains crystalline lipids then they may be prone to sedimen-
ing occurs (φc typically around 0.4 to 0.6 for a non-flocculated
tation because the density of lipids usually increases when they
conventional emulsion), the emulsion exhibits solid-like char-
crystallize. Oil-in-water nanoemulsions may also be prone to
acteristics, such as visco-elasticity and plasticity (Quemada and
sedimentation if they contain small oil droplets covered by a
Berli, 2002; Berli et al., 2005; McClements, 2005). For na-
relatively thick and dense shell (see below).
noemulsions, the oil concentration where the steep increase in
The velocity that a particle moves upwards in an emulsion or
viscosity occurs can be considerably less than that for conven-
nanoemulsion due to gravitational separation is given by Stokes’
tional emulsions because the interfacial shell contributes ap-
law:
preciably to the particles effective volume fraction (Fig. 17).
Experimental evidence of this effect is given in Fig. 18 (Weiss
et al., 2000), which shows that hydrocarbon oil-in-water emul-
2
2grparticle (ρparticle − ρ0 )
v=− (8)
sions undergo a transition from a liquid-like state when the 9η0
droplets are relatively large (d > 200 nm) to a gel-like state
when the droplets are relatively small (d < 80 nm). Similar where, v is the creaming velocity, rparticle is the particle radius,
results have also been reported for silicone oil-in-water emul- ρparticle is particle density, ρ0 is the continuous-phase density,
sions stabilized by an anionic surfactant (SDS), which can be η0 is the continuous-phase viscosity, and g is the acceleration
attributed to the “shell layer” around each oil droplet corre- due to gravity. For core-shell particles the radius is given by
sponding to the Debye screening length (Wilking and Mason, rparticle = rcore + δ, and the particle density depends on particle
2007). dimensions (rcore , δ) and the densities of the core and shell (see
The impact of the droplet characteristics on the overall rhe- below). Gravitational forces cause particles to move either up-
ology of an emulsion may be an important consideration when wards or downwards depending on their density relative to the
designing food products. Some products should have a rela- surrounding liquid. Hence, if only gravitational forces operated,
tively low viscosity (such as beverages) and so the droplets then the particles would accumulate at either the top or the bot-
present should not appreciably increase the overall viscosity. tom of a sample. In practice, particles may also move due to
Other products should be highly viscous or gel-like (e.g., dress- Brownian motion forces associated with the thermal energy of
ings, dips, deserts) and the droplets may play an important role the system. The root mean square distance moved by a√ particle
by thickening the system or forming a gel network. The above in a fluid due to Brownian motion is given by:  = (2Dt),
discussion suggests that it may be possible to prepare viscous where D is the translational diffusion coefficient of the particle
or gel-like nanoemulsions with rheological properties similar to and t is the time (Walstra, 2003). The dependence of the cream-
conventional emulsions of much higher fat content. ing velocity on the particle size for oil droplets (ρOil = 920 kg
m−3 ) suspended in water (ηWater = 1 mPa s, ρWater = 1000 kg
m−3 ) is shown in Fig. 19 (assuming no shell). The distance the
particles are predicted to move in one day due to gravitational
Stability forces and Brownian motion are plotted. The upward move-
ment of particles due to gravity increases with increasing par-
Like conventional emulsions, nanoemulsions are metastable ticle size for gravitational forces, but decreases with increasing
systems that tend to breakdown over time through a variety particle size for Brownian motion. Consequently, particle move-
of different physicochemical mechanisms, including gravita- ment is dominated by gravity for the relatively large particles in
tional separation, flocculation, coalescence, Ostwald ripening, emulsions, but dominated by Brownian motion for relatively
and chemical degradation (Dickinson, 1992; Friberg et al., 2004; small particles like those in some nanoemulsions. These calcu-
McClements, 2005). In this section, a brief overview of the im- lations indicate that we would not expect to see any creaming in
pact of droplet size on the physical and chemical stability of nanoemulsions when the particle diameter fell below about 70
nanoemulsions is given. nm due to the domination of Brownian motion effects.
312 D. J. McCLEMENTS AND J. RAO

Figure 17 Calculated dependence of the relative viscosity of oil-in-water nanoemulsions on mean particle diameter for systems with a shell thickness of 10 nm.

As mentioned earlier, the shell layer may form an appreciable shell makes little contribution to particle density). This indicates
fraction of the total volume of the droplets in nanoemulsions that it should be possible to produce density-matched particles
(Fig. 1). The overall density of a particle consisting of a core by controlling the oil-core size and the thickness of the adsorbed
(oil) surrounded by a shell layer (emulsifier) is given by: emulsifier layer.

ρparticle =
S ρS + (1 −
S )ρC (9) Droplet Aggregation
Nanoemulsions typically have much better stability to par-
The shell layer usually has a higher density than the oil or
ticle aggregation (flocculation and coalescence) than conven-
aqueous phases, so that an increase in the volume fraction of
tional emulsions because of the influence of their small particle
the shell layer will tend to increase the overall particle density.
size on colloidal interactions (Tadros et al., 2004). The nature of
This may have important implications in preventing gravita-
the colloidal interactions operating in a particular nanoemulsion
tional separation in emulsions since it will reduce the density
system depends on the bulk physicochemical properties of the
contrast between the particles and surrounding fluid (Fig. 20).
various phases (e.g., dielectric constant and refractive index),
In addition, very small particles may actually sediment rather
particle-core characteristics (such as radius), shell charac-
than cream if they contain sufficiently thick and dense emulsi-
teristics (such as thickness, charge, packing, rheology, and
fier layers. A practical example of this effect is shown in Fig.
hydrophobicity), and the properties of the intervening fluid (such
21, which compares the pH-stability of protein-stabilized con-
as pH, ionic strength, osmotic pressure, and temperature). To a
ventional emulsions (d43 = 325 nm) and nanoemulsions (d43 =
first approximation the overall colloidal interactions between a
66 nm) to gravitational separation. The small particles in the
pair of droplets in an emulsion or nanoemulsion can be described
nanoemulsions tend to sediment because the overall particle
by the sum of the van der Waals (wVDV ), electrostatic (wE ), steric
density is greater than water (δ/r is relatively large, so the shell
(wS ), and hydrophobic (wS ) interactions (McClements, 2005):
makes a large contribution to particle density), whereas the large
particles in conventional emulsions tend to cream because their
overall density is less than water (δ/r relatively small, so the w(h) = wVDV (h) + wE (h) + ws (h) + wH (h) (10)
FOOD-GRADE NANOEMULSIONS 313

Figure 18 Influence of particle radius on the rheology of SDS-stabilized oil-in-water nanoemulsions. The nanoemulsions gain solid like characteristics when
the droplet size falls below a critical level (Weiss and McClements, 2000).

The van der Waals and hydrophobic interactions are attrac- sions, which has been observed experimentally (Demetriades et
tive, whereas the steric and electrostatic interactions are usually al., 1997a; 1997b). A similar kind of behavior was observed in
repulsive. The steric interaction is a strong short-range repulsive the nanoemulsions, but the attractive and repulsive interactions
interaction, whereas the magnitude and range of the electrostatic were much weaker, and hence the height of the energy barrier
repulsion depends on the electrical charge on the droplets (ζ - and the depth of the primary minimum were much smaller (Fig.
potential) and the ionic composition of the aqueous phase (I). In 22b). In this case, one would expect there would be either little or
general, the magnitude of both the attractive and repulsive col- no droplet flocculation. This effect has recently been observed
loidal interactions tends to increase with increasing droplet size in protein-stabilized corn oil-in-water nanoemulsions (d43 =
(McClements, 2005). Theoretical predictions of the influence 60 nm) in our laboratory, where they were found to be stable to
of ionic strength on the overall interaction potential between droplet aggregation and creaming from 0 to 1000 mM NaCl (data
protein-coated oil droplets are shown in Fig. 22 for nanoemul- not shown). Thus both theoretical predictions and experimental
sions (r = 25 nm) and conventional emulsions (r = 250 nm). It measurements suggest that nanoemulsion delivery systems can
was assumed that the droplets had a relatively high ζ -potential be designed that have much better stability to droplet aggrega-
(+25 mV) so that there was a strong electrostatic repulsion be- tion and gravitational separation than conventional emulsions.
tween them in the absence of added salt. In the conventional
emulsions at low ionic strengths (I = 20 and 50 mM), there
was a high potential energy barrier (> 20 kT) due to electro- Ostwald Ripening
static repulsion that should prevent droplets from coming close
enough to aggregate (Fig. 22a). On the other hand, at high ionic Ostwald ripening is the process whereby the mean size of the
strengths (200 and 500 mM), this energy barrier was greatly droplets in an emulsion or nanoemulsion increases over time
diminished due to electrostatic screening effects and there was due to diffusion of oil molecules from small to large droplets
a deep primary minimum. The droplets would therefore be ex- through the intervening fluid (Kabalnov et al., 1992; Kabal-
pected to flocculate at high ionic strengths in conventional emul- nov, 2001). The driving force for this effect is the fact that the
314 D. J. McCLEMENTS AND J. RAO

Figure 19 Calculated influence of gravity and Brownian motion on the creaming velocity of differently sized oil droplets in oil–in–water emulsions.

water-solubility of an oil contained within a spherical droplet prepared using oils with a very low water-solubility, such as
increases as the size of the droplet decreases, which means that long-chain triglycerides (e.g., corn, soy, sunflower, or fish oils).
there is a higher concentration of solubilized oil molecules in On the other hand, OR may occur rapidly for nanoemulsions
the aqueous phase surrounding a small droplet than surrounding prepared using oils with an appreciable water-solubility, such as
a larger one (Kabalnov et al., 1992; McClements, 2005). Sol- short-chain triglycerides, flavor oils, and essential oils (Wooster
ubilized oil molecules therefore tend to move from around the et al., 2008; Li et al., 2009). Ostwald ripening can be retarded
smaller droplets to around the larger droplets because of this in these systems by adding a substance known as a ripening
concentration gradient. This leads to an increase in droplet size inhibitor. A ripening inhibitor is a non-polar molecule that is
over time, which can be described by the following equation soluble in the oil phase but insoluble in the water phase, e.g., a
once steady state has been achieved (Kabalnov et al., 1992): long-chain triacylglycerol. This type of molecule can inhibit OR
by generating entropy of mixing effect that counter-balances the
32 curvature effects. Consider an oil-in-water emulsion that con-
d(t)3 − d(0)3 = ωt = αS∞ Dt (11)
9 tains droplets comprised of two different lipid components—a
water-insoluble component and a water-soluble component. The
Here, d(t) is the number mean droplet diameter at time t, d0 water-soluble molecules will diffuse from the small to the large
is the initial number mean droplet diameter, ω is the Ostwald droplets due to OR. Consequently, there will be a greater con-
ripening rate, α = 2γ Vm /RT, S∞ is the water-solubility of the centration of water-insoluble molecules in the smaller droplets
oil phase in the aqueous phase, D is the translational diffusion than in the larger droplets after OR occurs. Differences in the
coefficient of the oil molecules through the aqueous phase, Vm is composition of emulsion droplets are thermodynamically un-
the molar volume of the oil, γ is the oil-water interfacial tension, favorable because of entropy of mixing—it is more favorable
R is the gas constant, and T is the absolute temperature. to have the two lipids distributed evenly throughout all of the
In practice, the main factor determining the stability of a na- droplets, rather than be located in particular droplets. Conse-
noemulsion to OR is the water-solubility of the oil phase (S∞ ). quently, there is a thermodynamic driving force that operates in
For this reason OR is not usually a problem for nanoemulsions the opposite direction to the OR effect. The change in droplet
FOOD-GRADE NANOEMULSIONS 315

Figure 20 Predicted influence of droplet size and shell layer thickness on the effective density of the core-shell particles in nanoemulsions. It was assumed that
the density of the shell was 1200 kg m−3 and that of the oil phase was 920 kg m−3 .

size distribution with time then depends on the concentration creases (A ∝ 1/r), which means that more of the lipid phase
and solubility of the two components within the oil droplets. will be exposed to the surrounding aqueous phase. If chemi-
This approach has previously been used to improve the stability cal degradation occurs primarily at the oil-water interface, then
of food-grade nanoemulsions, such as those containing short- reducing the lipid droplet size may promote the reaction rate.
chain triglycerides or essential oils (Wooster et al., 2008; Li et Second, if chemical degradation is accelerated by the presence
al., 2009). An example of this effect is shown in Fig. 23, which of UV or visible light, then a transparent nanoemulsion contain-
shows the change in droplet size during storage of orange oil- ing small droplets may be more susceptible to degradation than
in-water emulsions containing different levels of corn oil (the an opaque conventional emulsion because light can penetrate
ripening inhibitor). Orange oil (4-fold) has a relatively high sol- more easily. Nevertheless, one must also take into account any
ubility in water, and therefore is highly prone to OR, which leads differences in the properties of the interfacial layer surrounding
to an appreciable increase in mean droplet size during storage. the lipid droplets in nanoemulsions and emulsions.
On the other hand, corn oil has very low solubility in water, and
therefore it can retard OR if it is incorporated into the oil phase
prior to homogenization. Our results show that incorporating ≥ Molecular Distribution and Release Characteristics
10% corn oil into the lipid phase was sufficient for inhibiting
OR in these systems. The ability of nanoemulsions to incorporate and release func-
tional components, such as flavors, antimicrobials, vitamins,
nutraceuticals, and drugs, depends on the molecular character-
istics of the functional components as well as the composition,
Chemical Stability
properties, and microstructure of the nanoemulsions themselves
The small size of the lipid droplets in oil-in-water nanoemul- (McClements, 2005). Some encapsulated components are solu-
sions may also impact the chemical degradation rate of any ble in both oil and water phases and will move out of oil droplets
encapsulated components, e.g., due to oxidation or hydrolysis. when the emulsions are diluted with water. In principle, the re-
First, the oil-water surface area increases as the droplet size de- lease of this kind of encapsulated component can be delayed by
316 D. J. McCLEMENTS AND J. RAO

Figure 22b Predicted influence of changes in ionic strength on the total inter-
action potential between oil droplets coated by a globular protein: Nanoemulsion
(r = 25 nm, δ = 3 nm, ζ = 25 mV).

Figure 21 Influence of pH on the stability of protein-stabilized 0.5 wt% increasing the droplet size. However, the droplets in nanoemul-
corn oil-in-water conventional emulsions (r43 = 163 nm) and nanoemulsions
sions are so small that they have little effect on the release rate
(d43 = 33 nm) to gravitational separation. The particles in the nanoemulsions
moved downwards because their overall density was greater than water (due of encapsulated components through this mechanism. For en-
to the thick adsorbed protein layer), whereas the particles in the conventional capsulated components that have a very low water-solubility,
emulsions moved upwards because their density was less than water (Lee and one is typically interested in their release into the surrounding
McClements, 2010). aqueous phase in the human body, e.g., the mouth, stomach, or
small intestine. In this case, it is usually important to establish
any potential trigger mechanisms for release (e.g., dilution, pH,

Figure 22a Predicted influence of changes in ionic strength on the total inter- Figure 23 Dependence of mean droplet diameter on storage time for 5% oil-
action potential between oil droplets coated by a globular protein: Nanoemulsion in-water emulsions stabilized by modified starch on ratio of orange oil to corn
(r = 25 nm, δ = 3 nm, ζ = 25 mV). oil in the oil phase (pH 3.5, 20◦ C).
FOOD-GRADE NANOEMULSIONS 317

ionic strength, temperature, enzyme activity, etc.), as well as the or, to protect it from chemical degradation. One of the major ad-
rate and extent of release. In an emulsion-based delivery system, vantages of nanoemulsions over conventional emulsions for this
the release is usually characterized in terms of the increase in kind of application is that they can be incorporated into clear or
concentration of the encapsulated compound in the continuous slightly turbid products without altering their visual appearance.
phase or in some target organ (such as the mouth, nose, stom- (See section titled “Optical Properties.”)
ach, or gastrointestinal tract) as a function of time (Aguilera, Lipophilic components are usually blended with the oil phase
2006). The ability of nanoemulsions to encapsulate and release prior to nanoemulsion formation, so that they are trapped within
non-volatile components may be quite different from conven- the oil core of the droplets once the nanoemulsion has been
tional emulsions made from similar components. The solubility produced. The physical location of lipophilic components within
of a component within an emulsion droplet usually increases nanoemulsions depends on their molecular and physicochemi-
as its droplet size decreases. For a typical food oil contained cal properties, such as hydrophobicity, surface activity, oil-water
within a droplet surrounded by water (v = 10−3 m3 mol−1 , partition coefficient, solubility, and melting point. Highly non-
γ = 10 mJ m−2 ), the solubility increases 2.24-, 1.08-, 1.01-, polar lipophilic components tend to be distributed within the
and 1.00-fold for oil droplets with radii of 10, 100, 1000, and hydrophobic core of the particles, whereas more polar lipophilic
10,000 nm, respectively. This effect may alter the oil-water par- components may be incorporated within the amphiphilic shell.
tition coefficient (KOW ) of any substances encapsulated within The physical location of the lipophilic components may have
a nanoemulsion, and therefore their bioavailability. an important impact on the physical and chemical stability of
the system. For example, some lipophilic components are sus-
ceptible to chemical degradation when they come into contact
PRACTICAL APPLICATIONS OF NANOEMULSIONS
with particular water-soluble components in the aqueous phase.
In this case, it may be important to ensure that the lipophilic
Research into the development and application of food-grade components are trapped within the oil core, rather than within
nanoemulsions has increased recently because of the growing the shell. An example of this effect is the chemical degradation
need for effective delivery systems to encapsulate, protect and of citral molecules (a flavor oil component), which is promoted
release functional food components. Indeed, a recent report when these lipophilic components come into contact with pro-
(“Food Encapsulation: A Global Strategic Business Report,” tons in the aqueous phase (Choi et al., 2009; 2010; Mei et al.,
Global Industry Analysts, Inc., San Jose, CA) estimated that the 2010). In practice, it is often difficult to determine the precise
world market for food encapsulation will be nearly $40 billion location and orientation of lipophilic components within na-
by 2015. We have tabulated some commercial applications of noemulsion droplets. The analytical methods that can be used
nanoemulsions found on the Internet (Table 9). In the remainder are highly dependent on the nature of the lipophilic component
of this section, we provide an overview of potential applications used. Some of the methods that have proved useful for the in situ
of nanoemulsions in the food and beverage industries. determination of the location of lipophilic components within
emulsion droplets include nuclear magnetic resonance (Hamil-
Encapsulation of Lipophilic Components ton et al., 1996; Okamura et al., 2001; Nakaya et al., 2005;
Shen et al., 2005), fluorescence spectroscopy (Saito et al., 1996;
An important application of nanoemulsions in foods and bev- Tanaka et al., 2003) and Raman scattering microspectroscopy
erages is their ability to encapsulate lipophilic functional com- (Day et al., 2010).
ponents, such as vitamins, flavors, colors, antioxidants, preser- Nanoemulsions have been developed to encapsulate a num-
vatives, and nutraceuticals (Graves and Mason, 2008; Given, ber of different lipophilic components, such as β-carotene (Tan
2009). Encapsulation of a lipophilic component may be carried and Nakajima, 2005a,b; Chu et al., 2007a; 2008; Yuan et al.,
out for a number of reasons—to improve its ease of handling 2008a,b; Mao et al., 2009; Yin et al., 2009), citral (Choi et
and utilization; to facilitate its incorporation within a product; al., 2009; 2010; Mei et al., 2010), capsaicin (Choi et al., 2009),
to increase its bioavailability by improving the properties of flaxseed oil (Kentish et al., 2006), tributyrin (Li et al., 2009), co-
lipophilic compounds, such as solubility and efficient absorp- Enzyme Q (Ozaki et al., 2010), and oil-soluble vitamins (Relkin
tion through cells; to control the rate or location of its release; et al., 2009; Hatanaka et al., 2010).

Table 9 Some research articles on O/W nanoemulsions formed using the Emulsion Inversion Point method. Cremophor EL is a hydrogenated castor oil
surfactant
Oil Phase Surfactant/Co-surfactant Particle Diameter Reference

MCT Cremophor EL 28–78 nm Sadurn et al., 2005


Canola oil Span 20, Span 60, Span 80, Tween 20, Tween 60, Tween 80 215–233 nm de Morais et al., 2006
Paraffin oil Tween 20/Span 20 80–180 nm Pey et al., 2006
Hexadecane/styrene Brij 98 36 nm Sadtler et al., 2010
318 D. J. McCLEMENTS AND J. RAO

Increased Bioavailability and Controlled Delivery (Fig. 23). This can be achieved by preparing them from a carrier
of Lipophilic Components oil that has very low water-solubility (Wooster et al., 2008;
Li et al., 2009), or by controlling the mechanical properties
A number of studies have shown that the bioavailability of of the interfacial layer to restrict changes in droplet size (Mun
highly lipophilic functional components encapsulated within and McClements, 2006). In addition, the very small size of
lipid droplets is increased when the droplet size is decreased the droplets in nanoemulsions may promote the chemical
(Kesisoglou et al., 2007; Acosta, 2009). There are a number of degradation of encapsulated components (Mao et al., 2009).
possible reasons for this increase in bioavailability. (See section Nanoemulsions have very large specific surface areas so that
titled “Biological Fate, Bioavailability, and Potential Toxicity any chemical degradation reaction that occurs at the oil-water
of Nanoemulsions.”) interface may be promoted, such as lipid oxidation. In addition,
First, small droplets have a large surface area and may there- transparent nanoemulsions will allow UV and visible light to
fore be digested more quickly by digestive enzymes so that penetrate them easily, which may promote any light-sensitive
their contents are released and absorbed more easily. Second, chemical degradation reactions. Consequently, it may be nec-
the small droplets may penetrate into the mucous layer that essary to take additional steps to improve the chemical stability
coats the epithelium cells within the small intestine, thereby of labile components encapsulated within nanoemulsions, e.g.,
increasing their residence time and bringing them closer to the by adding antioxidants or chelating agents.
site of absorption. Third, very small particles may be directly
transported across the epithelium cell layer by either paracellu-
lar or transcellular mechanisms. Fourth, the water-solubility of Texture Modification
highly lipophilic components increases as the droplet size de-
creases, which may enhance absorption. At present there is no The fact that the droplets in nanoemulsions are very small
good understanding of the relative importance of these differ- means that the interfacial coating surrounding them may make
ent mechanisms for nanoemulsions with differing droplet sizes, an appreciable contribution to the particles overall volume
compositions, and surface characteristics. (Fig. 1). Consequently, it may be possible to promote the gela-
Recently, it has been shown that the oral availability of cur- tion of a nanoemulsion at a much lower oil concentration than
cumin can be increased by incorporating it within nanoemul- for a conventional emulsion (Fig. 18). This may be desirable
sions (Wang et al., 2008; Huang et al., 2010). Nanoemulsions for producing reduced fat products that need to be highly vis-
have also been shown to improve the bioavailability of various cous or gel like. At present, there is little work on utilizing
lipophilic nutraceuticals and pharmaceuticals (Hatanaka et al., nanoemulsions to create novel textural effects in edible prod-
2008; Nielsen et al., 2008; Ozaki et al., 2010; Talegaonkar et al., ucts. Nevertheless, recent studies with non-edible systems have
2010). It has also been proposed that nanoemulsions can be used shown that transparent nanoemulsions with gel-like characteris-
to effectively deliver nutraceuticals to specific sites within the tics can be prepared based on electrostatic repulsion effects, i.e.,
human body, which may lead to improvements of their efficacy when the thickness of the electrical double layer is on the or-
(Vishwanathan et al., 2009; Huang et al., 2010). For example, der of the radius of the emulsion droplets (Wilking and Mason,
the Salvona Company developed a delivery system (MultiSalTM ) 2007; Kawada et al., 2010). Potentially, this approach could be
to deliver multiple active ingredients (water-soluble and fat- used to produce highly viscous or visco-elastic nanoemulsions
soluble ingredients). It was proposed that this system enhances using much lower oil contents than is required to produce sim-
the stability and bioavailability of a wide range of nutrients and ilar systems in conventional emulsions (Wilking and Mason,
other ingredients, controls their release characteristics, and pro- 2007).
longs their residence time in the oral cavity, and thus prolongs
the sensation of flavors in the mouth. Recently, our laboratory
designed nanoemulsion-based delivery systems to control the di- Antimicrobial Activity
gestion and release of lipophilic food components (McClements
and Li, 2010). A number of studies suggest that nanoemulsions may be an
effective means of encapsulating and delivering antimicrobial
agents (Hamouda et al., 1999; 2001; Teixeira et al., 2007; Fer-
Improved Emulsion Stability reira et al., 2010). Antimicrobial nanoemulsions have been de-
veloped for the decontamination of food packaging equipment
As mentioned earlier, the stability of emulsions to grav- and for application to various food surfaces (Sekhon, 2010).
itational separation, flocculation, and coalescence is greatly Nanoemulsions based on non-ionic surfactants, soybean oil and
improved when their size is reduced, which may help extend the tributyl phosphate, have been reported to be highly effective
shelf-life of some emulsion-based products. (See section titled against a variety of food pathogens, including bacteria, viruses,
“Stability.”) Nevertheless, it is important to design the system fungi, and spores (Hamouda et al., 2001). However, they tend to
so that Ostwald ripening does not occur since nanoemulsions be much more effective against Gram-positive bacteria than
are highly susceptible to degradation through this mechanism against Gram-negative bacteria, which was attributed to the
FOOD-GRADE NANOEMULSIONS 319

influence of the cell-wall lipopolysaccharide that surrounds tosterols). The shell around nanoemulsion droplets may also
Gram-negative bacteria. The mechanisms by which antimicro- be comprised of one or more materials, including surfactants,
bial nanoemulsions inhibit microbial growth depend on the na- phospholipids, proteins, polysaccharides, and minerals, which
ture of the encapsulated antimicrobial agents they contain (e.g., may be digested to different rates and extents within the GI
essential oils, proteins, surfactants), as well as on the nature of tract. For example, phospholipids and proteins may be digested
the nanoemulsion droplets themselves (e.g., size, charge, com- by phospholipases and proteases in the stomach and small in-
position). Nanoemulsions can be designed to contain a variety testine, whereas dietary fibers may not be digested until they
of different antimicrobial agents that may act synergistically. reach the large intestine. The biological fate of a nanoemulsion
Nanoemulsion particles can be designed to deliver one or more particle may therefore depend on the nature of the core and shell
entrapped antimicrobial agents after they come into contact with material. At present there is a knowledge gap in understanding
the surfaces of microorganisms. The interaction between the how specific particle characteristics influence the biological fate
droplets and the microorganism can be enhanced by controlling of nanoemulsion particles within the GI tract, and it is clear that
the colloidal interactions in the system. For example, cationic more research is needed to establish their potential toxicity (Ha-
nanoemulsion particles can be made to bind strongly to anionic gens et al., 2007; Bouwmeester et al., 2009). In the remainder
bacteria surfaces through electrostatic attraction (Li et al., 2001), of this section, some of the potential ways that nanoemulsion
which may improve their antimicrobial efficacy by bringing the droplets could promote toxic effects due to their small size are
antimicrobial agents close to the site of action. Recently, it discussed.
has been shown that oil-in-water nanoemulsions prepared from
non-ionic surfactants and soybean oil are effective at inhibit-
ing bacterial growth in a variety of different biofilms (Teixeira Biological Fate of Nanoemulsion Droplets
et al., 2007).
An improved understanding of the potential toxic effects of
ingested nanoemulsions depends on knowledge of the physico-
BIOLOGICAL FATE, BIOAVAILABILITY, AND chemical and physiological processes that occur as nanoemul-
POTENTIAL TOXICITY OF NANOEMULSIONS sion droplets pass through the human body (Singh et al., 2009;
McClements et al., 2009).
There has been growing concern about the increased utiliza-
tion of nanoparticles in foods and beverages because of poten-
tial toxic effects associated with their small dimensions (Hagens Digestion
et al., 2007; Chaudhry et al., 2008; Bouwmeester et al., 2009; In this section, a brief overview of the complex conditions
Souto et al., 2009). Reducing the dimensions of a material to existing in different locations of the GI tract is given (Fig. 24),
the nanometer region may alter its biological fate within the and their potential influence on the properties of nanoemulsion
human body, such as the absorption, distribution, metabolism, droplets prior to absorption is discussed (McClements et al.,
and excretion (ADME) processes, thereby altering its potential 2009; Singh et al., 2009; van Aken, 2010b).
for promoting toxicity (Hagens et al., 2007; Bouwmeester et al., Mouth: After ingestion a nanoemulsion is mixed with saliva;
2009; Hu et al., 2009). The biological effects of nanoparticles de- changes in pH, ionic strength, and temperature; is acted upon
pend on their composition and physicochemical properties, such by digestive enzymes; interacts with the surfaces of the tongue,
as concentration, particle-size distribution, electrical charge, and mouth and throat; and, experiences a complex flow/force profile
interfacial characteristics. There is likely to be a large difference (van Aken, 2010a).
in the effects of non-digestible inorganic nanoparticles (such as
metals and minerals) compared to the effects of digestible or-
ganic nanoparticles (such as surfactants, lipids, proteins, and • As a result the nanoemulsion droplets may aggregate with
carbohydrates). The particles in nanoemulsions usually consist each other, or become coated with mucin molecules from the
of a core of lipophilic material surrounded by a shell of ad- saliva.
sorbed material (Fig. 1). As mentioned previously, the lipophilic • Stomach: After passing through the esophagus and entering
core may be comprised of one or more non-polar components, the stomach, the “bolus” is exposed to highly acidic condi-
including triacylglycerols, diacylglycerols, monoacylglycerols, tions (pH 1 to 3); mixed with enzymes (e.g., proteases and
flavor oils, essential oils, mineral oils, fat substitutes, weight- gastric lipases) and surface-active substances (e.g., phospho-
ing agents, oil-soluble vitamins, and nutraceuticals (such as lipids and proteins); and, experiences a complex flow/force
carotenoids, phytosterols, and Co-enzyme Q). Some of these profile (Kong and Singh, 2008; Ferrua and Singh, 2010).
lipophilic components are typically digested within the human Endogenous and exogenous surface-active substances may
gastrointestinal (GI) tract (e.g., triacylglycerols and diacylglyc- displace the original surface-active materials from the lipid
erols), some normally pass through the GI tract without being droplet surfaces through competitive adsorption processes.
absorbed (e.g., mineral oils and fat substitutes), while some Gastric lipases initiate lipid digestion, while gastric proteases
are absorbed without being digested (e.g., carotenoids and phy- initiate protein digestion.
320 D. J. McCLEMENTS AND J. RAO

Figure 24 Schematic diagram of the various physicochemical and physiological processes that may occur when nanoemulsions pass through the human GI
tract. The picture of the human body was obtained from http://en.wikipedia.org/wiki/Digestive tract (Copyright free).

• Small intestine: After entering the small intestine, the par- As a result of exposure to these various physiological con-
tially hydrolyzed and emulsified lipids are mixed with al- ditions, the droplets in nanoemulsions may change appreciably
kaline digestive juices containing bile salts, phospholipids, as they pass through the GI tract (Johnson, 2001; McClements
pancreatic lipase, colipase, and bicarbonate, which increase et al., 2007). For example, there may be changes in the com-
the pH close to neutral. Surface-active substances present in position, size distribution, aggregation state, electrical charge,
the small intestine compete with surface-active materials ini- interfacial properties, and physical state of the nanoemulsion
tially present at the lipid-droplet surfaces, and may displace droplets (Fig. 24). All of these changes need to be accounted for
them. Various digestion enzymes may hydrolyze the com- when trying to establish the potential biological fate and poten-
ponents present within a nanoemulsion droplet: pancreatic tial toxicity of nanoemulsion delivery systems. At present, there
lipase/co-lipase complex convert triglycerides and diglyc- is a poor understanding of the impact of specific physiological
erides into monoglycerides and free fatty acids; phospho- conditions on the fate of nanoemulsion droplets within the GI
lipases convert phospholipids to free fatty acids and other tract.
lipophilic substances; proteases convert proteins to peptides
and amino acids. Lipid digestion products are incorporated
within bile salt/phospholipid micelles and vesicles, and then Absorption
transported through the mucous layer to the surfaces of the Once the nanoemulsion droplets and/or their digestion prod-
enterocytes where they are absorbed (Fig. 24). The mucous ucts reach the surfaces of the GI tract they may be absorbed (De
layer has been reported to be about 30 to 100 µm thick (Roger Jong et al., 2008; Acosta, 2009; Hu et al., 2009). There are two
et al., 2010). Lipophilic components, such as nutraceuticals main types of cells where particle absorption may occur: epithe-
or pharmaceuticals, which are incorporated into the mixed lium cells and M-cells (des Rieux et al., 2006). Epithelium cells
micelles and vesicles, will also be transported to the surfaces are the most numerous type of cell lining the GI tract, but they
of the enterocytes. are not believed to be very efficient at absorbing particles. On
FOOD-GRADE NANOEMULSIONS 321

Figure 25 The droplets in nanoemulsions, and any functional components encapsulated within them, may be absorbed to a different extent than the droplets in
conventional emulsions within the human gastrointestinal tract.

the other hand, M-cells are believed to be highly efficient at ab- particle size for absorption in the nanometer range whose value
sorbing nanoparticles, but they are much less numerous (< 1% depends on particle type (Hu et al., 2009). Particle absorption
of epithelium surface) being mainly located in specialized re- is also believed to depend on the surface hydrophobicity of
gions called “Peyers patches.” Physiologically, the M-cells are nanoparticles (des Rieux et al., 2006). The physicochemical
responsible for absorbing various kinds of antigens, including, characteristics of a nanoparticles also determine their fate once
macromolecules, microorganisms, and certain types of particles, they have entered the epithelium cell: (i) They may be digested
which are then delivered to the underlying lymphoid system to by cellular enzymes into their constituent parts which may then
induce immune responses (des Rieux et al., 2006). Nanoparti- be absorbed; (ii) They may be transported directly out of the
cles may be absorbed by cells through a number of different cell and into the blood or lymph systems; or (iii) They may
mechanisms (Sahay et al., 2010a; 2010b): accumulate within specific locations in the cell (Hu et al., 2009;
Sahay et al., 2010a,b).
Some types of nanoparticles have been shown to promote
– Paracellular mechanism: Nanoparticles may be small enough
cellular damage if they accumulate within biological cells (Un-
to pass between the narrow gaps (“tight junctions”) separating
fried et al., 2007; Hu et al., 2009). The nanoparticles that
neighboring epithelial cells (Fig. 25). This mechanism has
are transported out of the epithelial cells via the portal vein
been suggested for the transport of solid lipid nanoparticles
or lymphatic system may circulate through the human body,
across epithelium cells (Yuan et al., 2007).
where they may then be metabolized, excreted, or accumulate
– Transcellular mechanism: Nanoparticles may be small
within certain tissues (Bouwmeester et al., 2009). As mentioned,
enough to be absorbed directly through epithelial cell mem-
these processes depend on the precise physicochemical charac-
branes by either passive or active transport mechanisms
teristics of the nanoparticles involved. For example, smaller
(Fig. 25). Typically, absorption occurs by an “endocytosis”
(metallic) particles have been found to accumulate within a
mechanism that involves the nanoparticle encountering the
wider range of tissues than larger particles (De Jong et al.,
cell membrane, the membrane wrapping itself around the
2008).
nanoparticle, and then part of the membrane budding-off to
One would expect that the direct absorption and accumu-
form a vesicle with a nanoparticle trapped inside.
lation of the oil droplets in edible nanoemulsions would be
unlikely because they would be expected to be rapidly digested
The pathway taken, and the rate and extent of nanoparticle within the stomach, small intestine, or epithelium cells. How-
uptake, depends on the properties of the nanoparticles, e.g., size, ever, if the nanoemulsion droplets were formed from indigestible
shape, charge, and interfacial chemistry (des Rieux et al., 2006; oils (such as hydrocarbons or mineral oils) or if the droplets were
Hu et al., 2009). Cationic nanoparticles have been shown to be coated with indigestible shells (such as dietary fibers) then di-
absorbed more readily than anionic nanoparticles (Hagens et al., rect absorption could occur. Direct particle absorption might
2007; Bouwmeester et al., 2009). Nanoparticles (d < 1000 nm) also occur if the nanoemulsion particles were able to diffuse
tend to be absorbed more readily by cells than microparticles through the mucous layer so rapidly that only limited digestion
(d > 1000 nm), but that there is often an optimum intermediate had time to occur. There is currently little empirical data on the
322 D. J. McCLEMENTS AND J. RAO

direct absorption of edible nanoemulsion particles by epithe- hydrolysis of digestible lipids (such as triacylglycerols), as
lium cells, and further work is clearly needed. As mentioned well as the extent of absorption of any encapsulated lipophilic
earlier, the physicochemical characteristics of the droplets in a substances. The particles in nanoemulsions may increase the
nanoemulsion may be altered considerably as they pass through transit time in the GI tract compared to those in conventional
the GI tract (e.g., their size, aggregation state, charge, and com- emulsions through a number of mechanisms (Bouwmeester
position), and these changes must be considered when studying et al., 2009):
particle absorption mechanisms. (i) The small size of the droplets in nanoemulsions means
they may be able to penetrate into the mucous layer
Metabolism, Distribution, and Excretion coating the enterocytes, whereas large droplets cannot
(Fig. 25);
If a nanoemulsion droplet is directly absorbed by the human
(ii) If the droplets in a nanoemulsion have a positive charge
body (rather than being digested first), then its distribution, ac-
in the small intestine they may also become trapped
cumulation, metabolism, and excretion may be different from
within the mucous layer (which consists of anionic
that of normally digested components (Hagens et al., 2007;
biopolymers).
Bouwmeester et al., 2009). Studies on metallic and inorganic
nanoparticles that are not digested in the GI tract have shown An increase in transit time should enable more of the lipids to
that they may accumulate in different organs, with the extent be digested, and a greater fraction of the bioactive component
of retention depending on their size (De Jong et al., 2008). For to be solubilized by micelles and absorbed by enterocytes
example, it was reported that smaller-sized gold nanoparticles (Acosta, 2009).
accumulated within a greater number of organs than larger-sized Digestion: Lipid digestion occurs at the oil-water interface sur-
ones. Further work is therefore needed to establish whether ed- rounding the emulsified lipid droplets and so the rate of lipid
ible nanoparticles can be directly absorbed by the human body, digestion tends to increase as the droplet size decreases, be-
and if they can, what their biological fate is. cause this increases the surface area of lipid exposed to the
digestive enzymes (Armand et al., 1992; Lundin et al., 2008).
Thus nanoemulsions should be digested more rapidly than
Biological Fate of Bioactive Components Encapsulated conventional emulsions, which may lead to the more rapid
within Nanoemulsion Droplets generation of lipid digestion products that can solubilize
bioactive components (see below), and to the faster release
Nanoemulsions are often used as delivery systems to en- of any encapsulated components. Nevertheless, the rate of
capsulate lipophilic components, such as ω-3 oils, carotenoids, lipid digestion may not always be faster for nanoemulsions
phytosterols, oil-soluble vitamins, and water-insoluble drugs than for conventional emulsions. Recent experiments in our
(McClements et al., 2007; Velikov et al., 2008; Acosta, 2009). laboratory found that globular protein-stabilized nanoemul-
It is therefore important to understand how the properties of sions prepared using the homogenization/solvent evapora-
nanoemulsions influence the release and absorption of these tion method were digested at a slower rate than conventional
encapsulated components. emulsions, which was attributed to the fact that the nanoemul-
The term bioavailability is defined as the fraction of an in- sions had a thicker protein coating than the conventional
gested component that eventually ends up in the systemic cir- emulsions.
culation (Versantvoort et al., 2004). The overall bioavailability The influence of the nature of the interfacial layer coating the oil
(F ) of lipophilic components that have poor water-solubility droplets on lipid digestion has also been shown in a number
depends on a number of physicochemical and physiological of other studies. If the interface is able to prevent the lipase
factors: molecules from adsorbing to the droplet surfaces, then the
rate and extent of lipid digestion may be retarded (Mun et al.,
F = FB × FA × FM (12) 2006; Chu et al., 2009). The composition of the carrier lipid
may also affect the rate and extent of digestion and absorp-
Here, FB is the fraction of lipophilic component released tion. Triacylglycerol oils (such as corn oil or fish oil) should
from the delivery system into the lumen of the gastrointesti- be fully digested within the stomach and small intestine,
nal tract to become bioaccessible, FA is the fraction of the whereas fat substitutes (such as Olestra) should not be. Thus,
released component that is absorbed across the intestinal ep- more of a bioactive may be released from a triacylglycerol oil
ithelia, and FM is the fraction of the absorbed component that than from a fat substitute. Lipid digestion also depends on the
reaches the systemic circulation without being metabolized. molecular characteristics of digestible triacylglycerols oils,
The characteristics of the droplets present within nanoemul- e.g., medium-chain triglycerides (MCT) tend to be digested
sions may affect these processes in a variety of ways (Acosta, more rapidly than long-chain triglycerides (corn oil) (Porter
2009): et al., 2007).
Solubilization: The particle characteristics within a nanoemul-
Transit Time: The time that a lipid particle spends within the sion may also influence the solubization of lipids and
GI tract may affect the rate and extent of the enzymatic lipophilic components within micelles and vesicles in the
FOOD-GRADE NANOEMULSIONS 323

small intestine (Porter et al., 2006; Porter et al., 2007). First, droplets in emulsions falls below a critical level within the 100–
the presence of the lipid carrier itself stimulates the release 1000 nm range (Acosta, 2009). For many bioactive components,
of bile salts and phospholipids from the bile duct, which an increase in bioavailability may be either desirable or may have
increases the number of micelles/vesicles present to solu- no adverse effects on human health. However, for certain bioac-
bilize the lipophilic component. Second, the surface-active tive components there may be concerns about increasing their
digestion products from the lipid carrier material, such as bioavailability, particularly those that exhibit toxic effects when
FFAs, MAGs, and DAGs, are incorporated into mixed mi- consumed at high levels. If one of these bioactive components
celles thereby increasing their solubilization capacity. normally has very low bioavailability but its absorption by the
Absorption: The relative bioavailability of highly lipophilic human body is increased substantially by incorporation into a
compounds has been shown to increase appreciably when nanoemulsion, then it could exhibit toxic effects that could not
the particle size is decreased below about 500 nm (Acosta, be predicted from data obtained on the same material in macro-
2009). The precise physicochemical mechanism underlying scopic form. This is particularly true if the bioactive component
this process is currently unknown, but a number of mecha- is incorporated into a product that is consumed regularly, such
nisms have been proposed (Fig. 25): as a soft drink or beverage emulsion.
(i) The water-solubility of highly lipophilic compounds in-
creases with decreasing particle size;
Direct Absorption of Nanoemulsion Droplets
(ii) Small particles can diffuse into the mucous layer coating
the enterocytes, which increases their retention time and As mentioned above, there is evidence that non-digestible
their ability to interact with the enterocytes; nanoparticles, such as metals (silver and gold) and inorganic
(iii) Very small particles may be able to cross the epithelia materials (titanium dioxide and silicon oxide), can cross the
by direct absorption into enterocytes through passive or epithelial layer by various mechanisms, e.g., paracellular or
active transcellular routes or by passing between cells transcellular (passive or active) (De Jong et al., 2008). Once ab-
by a paracellular route; sorbed into an epithelium cell, a nanoparticle may be digested,
(iv) Some lipophilic components are crystalline at room accumulate, or transported into the systemic circulation via the
temperature and have to be dissolved in carrier oil be- blood or lymph systems (Hu et al., 2009). The nanoparticles
fore homogenization. When the carrier oil is digested that are transported out of the epithelial cells may circulate
by lipase the lipophilic component may exceed its sol- through the human body, where they may then be metabolized,
ubility limit in the oil phase and form crystals. The size excreted, or accumulate within certain tissues (Bouwmeester
of the crystals formed will decrease as the size of the et al., 2009). These processes depend on the precise physic-
particles within nanoemulsions decreases, which may ochemical characteristics of the nanoparticles involved, e.g.,
have important consequences for the absorption of the composition, size, shape, charge, and interfacial chemistry. At
lipophilic component. present, there is no evidence that edible nanoparticles fabri-
Metabolism: If a nanoemulsion droplet is directly absorbed by cated from components that are normally digested within the
the human body (rather than being digested first) then the dis- GI tract are absorbed directly into the human body. Neverthe-
tribution, accumulation, metabolism, and excretion of any en- less, if the droplets in a nanoemulsion were prepared using an
capsulated components may be altered (Hagens et al., 2007; indigestible oil (such as hydrocarbons or mineral oils) or if the
Bouwmeester et al., 2009). droplets were coated with an indigestible shell (such as dietary
fibers) then direct absorption could occur. Hence, further work
is needed to determine whether this mechanism is important in
Potential Toxicity of Nanoemulsions humans.

At present there is little experimental evidence on the poten-


tial toxicity of food-grade nanoemulsions. Nevertheless, there Interference with Normal GI Function
are a number of physicochemical and physiological mechanisms Nanoparticles could alter the normal function of the GI tract
associated with the small particle size in nanoemulsions that as they pass through the mouth, stomach, and small intestine
could potentially cause toxicity. Currently, there is no standard- due to their small size, which could lead to some adverse effects
ized testing protocol specifically designed to assess the potential (Chaudhry et al., 2008). For example, it may be possible for very
toxicity of nanoemulsions intended for utilization within food small droplets to absorb directly through epithelium cells in the
products (Maynard et al., 2006). An outline of a potential ap- mouth, esophagus, or stomach before being digested within the
proach to evaluate the toxicity of nanoemulsions is highlighted small intestine. Very small droplets have a high specific surface
in Fig. 26. area and high curvature, and different surface reactivity than
bulk materials, which may alter the accumulation and activ-
Increased Bioavailability of Bioactive Components that are
ity of bile salts, lipase, and other digestive components at the
Toxic at High Levels
droplet surfaces, thereby altering the normal GI function. For
As mentioned previously, the bioavailability of encapsulated example, adsorption of proteins to particle surfaces, may lead
lipophilic components often increases when the size of the to denaturation and loss of normal function, which could have
324 D. J. McCLEMENTS AND J. RAO

Figure 26 Proposed protocol for testing the potential toxicity of food-grade nanoemulsions.

adverse effects on human health (Hoet et al., 2004). The particles involves the utilization of organic solvents, such as acetone,
in nanoemulsions may be able to attach to cellular membrane hexane, or ethyl acetate (Horn and Rieger, 2001). These organic
receptors, disrupting the cell’s normal metabolism and function. solvents are usually removed by evaporation during the prepara-
In conclusion, the combination of small size, high-surface area, tion of the nanoemulsion, but some residual solvent may remain
and high-surface energy for the particles in nanoemulsions may in the final product. It is therefore important to be aware of
lead to effects in biological systems that are not predictable from the potential toxic effects associated with any residual organic
the bulk form of the same materials (Jiang et al., 2009). solvents if nanoemulsions are produced using this approach.
The potential toxicity of emulsifiers and solvents that are
suitable for utilization within foods, pharmaceuticals, and other
Compositional Effects consumer products have been published by various organiza-
tions, including the World Health Organization (www.who.int),
Some of the components typically used to prepare nanoemul- the Food and Drug Administration (www.fda.gov) and the Eu-
sions are toxic when consumed at sufficiently high levels, e.g., ropean Food Safety Authority (www.efsa.europa.eu). These or-
emulsifiers and solvents. A relatively large amount of emulsi- ganizations carry out technical reviews of the toxicity data of
fier is required to cover the large specific surface area associated different additives and establish safe usage levels and analytical
with the droplets in nanoemulsions. At present, the most widely protocols for testing toxicity.
used emulsifiers for the preparation of nanoemulsions are small
molecule surfactants (and sometimes co-surfactants). This is
mainly because of the ability of surfactants to spontaneously
form nanoemulsions by low-energy methods (such as the PIT CONCLUSIONS
method), and because of their ability to rapidly adsorb to droplet
surfaces and reduce the interfacial tension in high-energy meth- Conventional oil-in-water emulsions are currently the
ods (such as high-pressure homogenization). Natural emulsi- most widely used emulsion-based delivery systems in many
fiers, such as proteins, polysaccharides, and phospholipids, tend industrial applications. These emulsions contain droplets that
to be much less effective at forming nanoemulsions. Many small are relatively large (r > 100 nm), and so they are optically
molecule surfactants are known to exhibit toxic effects when opaque and susceptible to breakdown through gravitational
consumed at high levels (Kralova et al., 2009; He et al., 2010). separation and droplet aggregation mechanisms. In contrast,
Consequently, the higher usage of surfactants in nanoemulsions nanoemulsions contain relatively small droplets (r < 100 nm)
compared to conventional emulsions could lead to some adverse that do not scatter light strongly and that are highly stable to
health effects. gravitational separation and droplet aggregation. In addition,
The preparation of nanoemulsions using some low-energy the small size of the droplets in nanoemulsions means that they
methods (e.g., solvent displacement or evaporation methods) are able to greatly increase the bioavailability of encapsulated
FOOD-GRADE NANOEMULSIONS 325

lipophilic substances. Nanoemulsions may therefore have ap- Aguilera, J.M. (2006). Food microstructure affects the bioavailability of several
plications in the food, beverage, and pharmaceutical industries nutrients. J. Food Sci. 72(2):R21–R32.
Amani, A., York, P., Chrystyn, H., and B. Clark (2010). Factors affecting the
as delivery systems, particularly in products that need to be
stability of nanoemulsions¡a Use of artificial neural networks. Pharm. Res.
optically transparent, or when increased bioavailability of an 27(1):37–45.
active component is important. Anton, N., Benoit, J.P., and P. Saulnier (2008). Design and production of
nanoparticles formulated from nano-emulsion templates : A review. Jour-
nal of Controlled Release 128():185–199.
FUTURE TRENDS Anton, N., Gayet, J.P., Benoit, P., and P. Saulnier (2007). Nano-emulsions
and nanocapsules by the pit method: An investigation on the role of the
temperature cycling on the emulsion phase inversion. International Journal
Nanoemulsions are likely to become increasingly the focus of of Pharmaceutics 344(1–2):44–52.
research and development efforts because of their potential ad- Anton, N., and T.F. Vandamme (2009). The universality of low-energy nano-
vantages over conventional emulsions for certain applications, emulsification. International Journal of Pharmaceutics 377(1–2):142–147.
Armand, M., Borel, P., Ythier, P., Dutot, G., Melin, C., Senft, M., H. Lafont, and
e.g., transparent foods and beverages, increased bioavailability,
D. Lairon (1992). Effects of droplet size, triacylglycerol composition, and
and improved physical stability. Nevertheless, there are a num- calcium on the hydrolysis of complex emulsions by pancreatic lipase: An in
ber of challenges that need to be overcome before nanoemul- vitro study. Journal of Nutritional Biochemistry 3(7):333–341.
sions are more widely used. First, suitable food-grade ingre- Asker, D., J. Weiss, and McClements, D.J. (2009). Analysis of the interactions
dients must be identified for formulating food nanoemulsions. of a cationic surfactant (lauric arginate) with an anionic biopolymer (pectin):
Isothermal titration calorimetry, light scattering, and microelectrophoresis.
Currently, many of the components used to fabricate nanoemul-
Langmuir 25(1):116–122.
sions using both low-energy and high-energy approaches are Berli, C.L.A., Deiber, J.A., and Quemada, D. (2005). On the viscosity of con-
unsuitable for widespread utilization within the food industry, centrated suspensions of charged colloids. Latin American Applied Research
e.g., synthetic surfactants, synthetic polymers, synthetic oils, or 35(1):15–22.
organic solvents. Ideally, the food industry would like to prepare Bos, M.A., and van, T. Vliet (2001). Interfacial rheological properties of ad-
sorbed protein layers and surfactants: A review. Advances in Colloid and
nanoemulsions from commonly used and acceptable food grade
Interface Science 91():437–471.
ingredients (such as flavor oils, triglyceride oils, proteins, and Bouchemal, K., Briançon, S., E. Perrier, and Fessi, H. (2004). Nano-emulsion
polysaccharides). Second, suitable processing operations must formulation using spontaneous emulsification: Solvent, oil and surfac-
be identified to economically and robustly fabricate food-grade tant optimisation. International Journal of Pharmaceutics. 280(1–2):241–
nanoemulsions on an industrial scale. Many of the approaches 251.
Bouton, F., Durand, M., V. Nardello-Rataj, M. Serry, and Aubry, J.M. (2009).
that have been developed within research laboratories may not
Classification of terpene oils using the fish diagrams and the equivalent alkane
be suitable for scale-up to industrial production facilities. In ad- carbon (Eacn) scale. Colloids and Surfaces A-Physicochemical and Engineer-
dition, it would be advantageous to utilize existing equipment ing Aspects. 338(1–3):142–147.
and manufacturing lines, so as to reduce the costs of develop- Bouwmeester, H., Dekkers, S., Noordam, M.Y., Hagens, W.I., Bulder, A.S., C. de
ing new manufacturing processes. Third, there may be some Heer, S. ten Voorde, S.W.P. Wijnhoven., H.J.P. Marvin, and Sips, A. (2009).
Review of health safety aspects of nanotechnologies in food production.
safety concerns associated with the utilization of very small
Regulatory Toxicology and Pharmacology, 53(1):52–62.
lipid droplets in foods, e.g., they may change the extent or route Bummer, P.M. (2004). Physical chemical considerations of lipid-based oral
of absorption of lipophilic components. Thus, the bioavailabil- drug delivery-solid lipid nanoparticles. Critical Reviews in Therapeutic Drug
ity or potential toxicity of a lipophilic component encapsulated Carrier Systems, 21(1):1–20.
within nanometer-sized lipid droplets may be considerably dif- Capek, I. (2004). Degradation of kinetically-stable O/W emulsions. Advances
in Colloid and Interface Science, 107(2–3):125–155.
ferent from when it is dispersed within a bulk lipid phase.
Chanamai, R., and McClements, D.J. (2002). Comparison of gum arabic, mod-
ified starch, and whey protein isolate as emulsifiers: Influence of ph, CaCl2
and temperature. Journal of Food Science, 67(1):120–125.
ACKNOWLEDGEMENTS Chantrapornchai, W., Clydesdale, F., and McClements, D.J. (1998). Influence of
droplet size and concentration on the color of oil-in-water emulsions. Journal
of Agricultural and Food Chemistry, 46(8):2914–2920.
This material is based upon work supported by the Coop- Chantrapornchai, W., Clydesdale, F., and McClements, D.J. (1999). Influence of
erative State Research, Extension, Education Service, United droplet characteristics on the optical properties of colored oil-in-water emul-
States Department of Agriculture, Massachusetts Agricultural sions. Colloids and Surfaces a-Physicochemical and Engineering Aspects,
Experiment Station, and United States Department of Agricul- 155(2–3):373–382.
ture, NIFA Grant. Chantrapornchai, W., Clydesdale, F.M., and McClements, D.J. (2001). Influence
of relative refractive index on optical properties of emulsions. Food Research
International, 34(9):827–835.
Chaudhry, Q., Scotter, M., Blackburn, J., Ross, B., Boxall, A., Castle, L., Aitken,
REFERENCES R., and Watkins, R. (2008). Applications and implications of nanotechnolo-
gies for the food sector. Food Additives and Contaminants, 25(): 241–258.
Abismail, B., Canselier, A., Wilhelm, M., Delmas, H., and C. Gourdon (1999). Chiu, W. (2006). The Preparation and Physical Properties of Jojoba Oil Nano-
Emulsification by ultrasound: Drop size distribution and stability. Ultrasonics Emulsion, Doctoral Dissertation, Department of Applied Chemistry, China.
Sonochemistry 6(1–2):75–83. Choi, A.J., Kim, C.J., Cho, Y.J., Hwang, J.K., and Kim, C.T. (2009). Effects of
Acosta, E. (2009). Bioavailability of nanoparticles in nutrient and nutraceutical surfactants on the formation and stability of capsaicin-loaded nanoemulsions.
delivery. Current Opinion in Colloid & Interface Science 14(1):3–15. Food Science and Biotechnology, 18(5):1161–1172.
326 D. J. McCLEMENTS AND J. RAO

Choi, S.J., Decker, E.A., Henson, L., Popplewell, L.M., and McClements, D.J. Flanagan, J., and Singh, H. (2006). Microemulsions: A potential delivery sys-
(2009). Stability of citral in oil-in-water emulsions prepared with medium- tem for bioactives in food. Critical Reviews in Food Science and Nutrition,
chain triacylglycerols and triacetin. Journal of Agricultural and Food Chem- 46():221–237.
istry, 57(23):11349–11353. Freitas, S., Merkle, H.P., and Gander, B. (2005). Microencapsulation by solvent
Choi, S.J., Decker, E.A., Henson, L., Popplewell, L.M., and McClements, D.J. extraction/evaporation: Reviewing the state of the art of microsphere prepa-
(2010). Inhibition of citral degradation in model beverage emulsions using ration process technology. Journal of Controlled Release, 102(2):313–332.
micelles and reverse micelles. Food Chemistry, 122(1):111–116. Friberg, S., Larsson, K., and Sjoblom, J. (2004). Food Emulsions. Marcel
Chu, B., Ichikawa, S., Kanafusa, S., and Nakajima, M. (2007a). Preparation of Dekker, New York.
protein-stabilized -carotene nanodispersions by emulsification–evaporation Garti, N., Yaghmur, A., Leser, M., Clement, V., and Watzke, H. (2001). Improved
method. Journal of the American Oil Chemists’ Society, 84(11):1053–1062. oil solubilization in oil/water food grade microemulsions in the presence of
Chu, B.S., Ichikawa, S., Kanafusa, S., and Nakajima, M. (2007b). Prepara- polyols and ethanol. Journal of Agricultural and Food Chemistry, 49(5):2552–
tion and characterization of beta-carotene nanodispersions prepared by sol- 2562.
vent displacement technique. Journal of Agricultural and Food Chemistry, Genovese, D.B., Lozano, J.E., and Rao, M.A. (2007). The rheology of colloidal
55(16):6754–6760. and Noncolloidal food dispersions. Journal of Food Science, 72(2):R11–R20.
Chu, B.S., Ichikawa, S., Kanafusa, S., and Nakajima, M. (2008). Stability of Gibout, S., Jamil, A., Kousksou, T., Zeraouli, Y., and Castaing, J.-Lasvignottes
protein-stabilised beta-carotene nanodispersions against heating, salts and ph. (2007). Experimental determination of the nucleation probability in emul-
Journal of the Science of Food and Agriculture, 88(10):1764–1769. sions. Thermochimica Acta, 454(1):57–63.
Chu, B.S., Rich, G.T., Ridout, M.J., Faulks, R.M., M.S.J. Wickham, and Wilde, Given, P.S. (2009). Encapsulation of flavors in emulsions for beverages. Current
P.J. (2009). Modulating pancreatic lipase activity with galactolipids: Effects Opinion in Colloid & Interface Science, 14(1):43–47.
of emulsion interfacial composition. Langmuir, 25(16):9352–9360. Gradzielski, M. (1998). Effect of the cosurfactant structure on the bending
Cooper, C.L., Dubin, P.L., Kayitmazer, A.B., and Turksen, S. (2005). elasticity in nonionic oil-in-water microemulsions. Langmuir, 14(21):6037–
Polyelectrolyte-protein complexes. Current Opinion in Colloid & Interface 6044.
Science, 10(1–2):52–78. Graves, S.M., and Mason, T.G. (2008). Transmission of visible and ultraviolet
Dalgleish, D.G., West, S.J., and Hallett, F.R. (1997). The characterization of light through charge-stabilized nanoemulsions. Journal of Physical Chemistry
small emulsion droplets made from milk proteins and triglyceride oil. Col- C, 112(33):12669–12676.
loids and Surfaces A-Physicochemical and Engineering Aspects, 123:145– Gu, Y.S., Decker, E.A., and McClements, D.J. (2005). Influence of ph and
153. carrageenan type on properties of beta-lactoglobulin stabilized oil-in-water
Day, J.P.R., Rago, G., Domke, K.F., Velikov, K.P., and Bonn, M. (2010). Label- emulsions. Food Hydrocolloids, 19(1):83–91.
free imaging of lipophilic bioactive molecules during lipid digestion by mul- Gutierrez, J.M., Gonzalez, C., Maestro, A., Sole, I., Pey, C.M., and Nolla, J.
tiplex coherent anti-stokes Raman scattering microspectroscopy. Journal of (2008). Nano-emulsions: New applications and optimization of their prepa-
the American Chemical Society, 132(24):8433–8439. ration. Current Opinion in Colloid & Interface Science, 13(4):245–251.
De Jong, W.H., Hagens, W.I., Krystek, P., Burger, M.C., Sips, A., and Geertsma, Guzey, D., and McClements, D.J. (2006a). Formation, stability and properties
R.E. (2008). Particle size-dependent organ distribution of gold nanoparticles of multilayer emulsions for application in the food industry. Advances in
after intravenous administration. Biomaterials, 29(12):1912–1919. Colloid and Interface Science, 128:227–248.
de Morais, J., Santos, O., Delicato, T., Goncalves, R., and Rocha, P.-Filho Guzey, D., and McClements, D.J. (2006b). Influence of environmental stresses
(2006). Physicochemical characterization of canola oil/water nano-emulsions on o/w emulsions stabilized by β-lactoglobulin–pectin and β-lactoglobulin–
obtained by determination of required hlb number and emulsion phase in- pectin–chitosan membranes produced by the electrostatic layer-by-layer de-
version methods. Journal of Dispersion Science and Technology, 27(1): position technique. Food Biophysics, 1(1):30–40.
109–115. Hagens, W.I., Oomen, A.G., W.H. de Jong, Cassee, F.R., and Sips, A. (2007).
Demetriades, K., Coupland, J.N., and McClements, D.J. (1997a). Physical prop- What do we (need to) know about the kinetic properties of nanoparticles in
erties of whey protein stabilized emulsions as related to ph and nacl. Journal the body? Regulatory Toxicology and Pharmacology, 49():217–229.
of Food Science, 62(2):342–347. Hamilton, J.A., Vural, J.M., Carpentier, Y.A., and Deckelbaum, R.J. (1996).
Demetriades, K., Coupland, J.N., and McClements, D.J. (1997b). Physicochem- Incorporation of medium chain triacylglycerols into phospholipid bilayers:
ical properties of whey protein-stabilized emulsions as affected by heating effect of long chain triacylglycerols, cholesterol, and cholesteryl esters. Jour-
and ionic strength. Journal of Food Science, 62():462–467. nal of Lipid Research, 37(4):773–782.
des Rieux, A., Fievez, V., Garinot, M., Schneider, Y.J., and Preat, V. (2006). Hamouda, T., Hayes, M.M., Cao, Z.Y., Tonda, R., Johnson, K., Wright, D.C.,
Nanoparticles as potential oral delivery systems of proteins and vaccines: A Brisker, J., and Baker, J.R. (1999). A novel surfactant nanoemulsion with
mechanistic approach. Journal of Controlled Release, 116(1):1–27. broad-spectrum sporicidal activity against bacillus species. Journal of Infec-
Dickinson, E. (1992). Introduction to Food Colloids. Royal Society of Chem- tious Diseases, 180(6):1939–1949.
istry, Cambridge, UK. Hamouda, T., Myc, A., Donovan, B., Shih, A.Y., Reuter, J.D., and Baker, J.R.
Dickinson, E. (1997). Enzymic crosslinking as a tool for food colloid rheology (2001). A novel surfactant nanoemulsion with a unique non-irritant topical
control and interfacial stabilization. Trends in Food Science & Technology, antimicrobial activity against bacteria, enveloped viruses and fungi. Microbi-
8(10):334–339. ological Research, 156(1):1–7.
Dickinson, E. (2003). Hydrocolloids at interfaces and the influence on the Hashida, M., Kawakami, S., and Yamashita, F. (2005). Lipid carrier systems
properties of dispersed systems. Food Hydrocolloids, 17(1):25–39. for targeted drug and gene delivery. Chemical & Pharmaceutical Bulletin,
Do, L., Withayyapayanon, A., Harwell, J., and Sabatini, D. (2009). Environ- 53(8):871–880.
mentally friendly vegetable oil microemulsions using extended surfactants Hatanaka, J., Chikamori, H., Sato, H., Uchida, S., Debari, K., Onoue, S., and
and linkers. Journal of Surfactants and Detergents, 12(2):91–99. Yamada, S. (2010). Physicochemical and pharmacological characterization
Fernandez, P., Andre, V., Rieger, J., and Kuhnle, A. (2004). Nano- of alpha-tocopherol-loaded nano-emulsion system. International Journal of
emulsion formation by emulsion phase inversion. Colloids and Surfaces Pharmaceutics, 396(1–2):188–193.
A-Physicochemical and Engineering Aspects, 251(1–3):53–58. Hatanaka, J., Kimura, Y., Z. Lai-Fu, Onoue, S., and Yamada, S. (2008). Physic-
Ferreira, J.P., Alves, D., Neves, O., Silva, J., Gibbs, P.A., and Teixeira, P.C. ochemical and pharmacokinetic characterization of water-soluble coenzyme
(2010). Effects of the components of two antimicrobial emulsions on food- q(10) formulations. International Journal of Pharmaceutics, 363(1–2):112–
borne pathogens. Food Control, 21():227–230. 117.
Ferrua, M.J., and Singh, R.P. (2010). Modeling the fluid dynamics in a hu- He, C.X., He, Z.G., and Gao, J.Q. (2010). Microemulsions as drug delivery
man stomach to gain insight of food digestion. Journal of Food Science, systems to improve the solubility and the bioavailability of poorly water-
75(7):R151-R162. soluble drugs. Expert Opinion on Drug Delivery, 7(4):445–460.
FOOD-GRADE NANOEMULSIONS 327

Henry, J., Fryer, P., Frith, W., and Norton, I. (2009). emulsification mechanism Innovative Foods Centre Conference on Food Innovation, Emerging Science,
and storage instabilities of hydrocarbon-in-water sub-micron emulsions sta- Technologies and Applications, Melbourne, Australia.
bilised with tweens (20 and 80), brij 96v and sucrose monoesters. Journal of Kesisoglou, F., Panmai, S., and Wu, Y.H. (2007). Application of nanoparticles
Colloid and Interface Science, 338(1):201–206. in oral delivery of immediate release formulations. Current Nanoscience,
Henry, J.V.L., Fryer, P.J., Frith, W.J., and Norton, I.T. (2010). The influence of 3(2):183–190.
phospholipids and food proteins on the size and stability of model sub-micron Klein, M., Aserin, A., Svitov, I., and Garti, N. (2010). Enhanced stabilization
emulsions. Food Hydrocolloids, 24(1):66–71. of cloudy emulsions with gum arabic and whey protein isolate. Colloids and
Hoeller, S., Sperger, A., and Valenta, C. (2009). Lecithin based nanoemulsions: Surfaces B-Biointerfaces, 77(1):75–81.
a Comparative study of the influence of non-ionic surfactants and the cationic Kong, F., and Singh, R.P. (2008). Disintegration of solid foods in human stom-
phytosphingosine on physicochemical behaviour and skin permeation. Inter- ach. Journal of Food Science, 73(5):R67–R80.
national Journal of Pharmaceutics, 370(1–2):181–186. Kralova, I., and Sjoblom, J. (2009). Surfactants used in food industry: A review.
Hoet, P., I. Brüske-Hohlfeld, and Salata, O. (2004). Nanoparticles – known and Journal of Dispersion Science and Technology, 30(9):1363–1383.
unknown health risks. Journal of Nanobiotechnology, 2(1):12–26. Leal-Calderon, F., Schmitt, V., and Bibette, J. (2007). Emulsion Science: Basic
Horn, D., and Rieger, J. (2001). Organic nanoparticles in the aqueous phase Principles. Springer Verlag: Berlin, Germany.
– theory, experiment, and use. Angewandte Chemie-International Edition, Lee, S.J., and McClements, D.J. (2010). Fabrication of protein-stabilized na-
40(23):4331–4361. noemulsions using a combined homogenization and amphiphilic solvent dis-
Howard, M.D., Jay, M., Dziublal, T.D., and Lu, X.L. (2008). Pegylation of solution/evaporation approach. Food Hydrocolloids, 24(6–7):560–569.
nanocarrier drug delivery systems: State of the art. Journal of Biomedical Leong, T., Wooster, T., Kentish, S., and Ashokkumar, M. (2009). Minimising
Nanotechnology, 4(2):133–148. oil droplet size using ultrasonic emulsification. Ultrasonics Sonochemistry,
Hu, L., Mao, Z.W., and Gao, C.Y. (2009). Colloidal particles for cellular uptake 16(6):721–727.
and delivery. Journal of Materials Chemistry, 19(20):3108–3115. Leong, T.S.H., Wooster, T.J., Kentish, S.E., and Ashokkumar, M. (2009). Min-
Huang, Q.R., Yu, H.L., and Ru, Q.M. (2010). Bioavailability and delivery of imising oil droplet size using ultrasonic emulsification. Ultrasonics Sono-
nutraceuticals using nanotechnology. Journal of Food Science, 75(1):R50- chemistry, 16(6):721–727.
R57. Li, J., D.J. McClements, and McLandsborough, L.A. (2001). Interaction be-
Hunter, R.J. (1986). Foundations of Colloid Science. Oxford University Press, tween emulsion droplets and escherichia coli cells. Journal of Food Science,
Oxford, UK. 66(4):570–574.
Imeson, A. (2010). Food Stabilizers, Thickeners and Gelling Agents. John Wiley Li, Y., S. Le Maux, Xiao, H., and McClements, D.J. (2009). Emulsion-based
& Sons, Chichester, UK delivery systems for tributyrin, a potential colon cancer preventative agent.
Israelachvili, J. (1992). Intermolecular and Surface Forces, Second Edition. Journal of Agricultural and Food Chemistry, 57(19):9243–9249.
Academic Press, London. Lin, C.Y., and Chen, L.W. (2008). Comparison of fuel properties and emis-
Jafari, S., He, Y., and Bhandari, B. (2007). Optimization of nano-emulsions sion characteristics of two- and three-phase emulsions prepared by ultra-
production by microfluidization. European Food Research and Technology, sonically vibrating and mechanically homogenizing emulsification methods.
225(5):733–741. Fuel, 87(10–11):2154–2161.
Jafari, S.M., Assadpoor, E., He, Y.H., and Bhandari, B. (2008). Re-coalescence Littoz, F., and McClements, D.J. (2008). Bio-mimetic approach to improving
of emulsion droplets during high-energy emulsification. Food Hydrocolloids, emulsion stability: Cross-linking adsorbed beet pectin layers using laccase.
22(7):1191–1202. Food Hydrocolloids, 22(7):1203–1211.
Jafari, S.M., He, Y.H., and Bhandari, B. (2006). Nano-emulsion production Liu, W., Sun, D., Li, C., Liu, Q., and Xu, J. (2006). Formation and stabil-
by sonication and microfluidization: A comparison. International Journal of ity of paraffin oil-in-water nano-emulsions prepared by the emulsion inver-
Food Properties, 9():475–485. sion point method. Journal of Colloid and Interface Science, 303(2):557–
Jiang, J., G. Oberdörster, and Biswas, P. (2009). Characterization of size, surface 563.
charge, and agglomeration state of nanoparticle dispersions for toxicological Lundin, L., Golding, M., and Wooster, T.J. (2008). Understanding food structure
studies. Journal of Nanoparticle Research, 11(1):77–89. and function in developing food for appetite control. Nutrition & Dietetics,
Johnson, L.R. (2001). Gastrointestinal Physiology, 6th Edition. Mosby, St. 65:S79–S85.
Louis, Missouri. Maa, Y.F., and Hsu, C.C. (1999). Performance of sonication and microflu-
Johnston, A.P.R., Cortez, C., Angelatos, A.S., and Caruso, F. (2006). Layer-by- idization for liquid-liquid emulsification. Pharmaceutical Development and
layer engineered capsules and their applications. Current Opinion in Colloid Technology, 4(2):233–240.
& Interface Science, 11(4):203–209. Maestro, A., Sole, I., Gonzalez, C., Solans, C., and Gutierrez, J.M. (2008).
Kabalnov, A. (2001). Ostwald ripening and related phenomena. Journal of Influence of the phase behavior on the properties of ionic nanoemulsions
Dispersion Science and Technology, 22(1):1–12. prepared by the phase inversion composition method. Journal of Colloid and
Kabalnov, A.S., Pertzov, A.V., and Shchukin, E.D. (1987). Ostwald ripening Interface Science, 327(2):433–439.
in 2-component Disperse phase systems: Application to emulsion stability. Mao, L.K., Xu, D., X., Yang, J., Yuan, F., Gao, Y.X., and Zhao, J. (2009). Effects
Colloids and Surfaces, 24(1):19–32. of small and large molecule emulsifiers on the characteristics of beta-carotene
Kabalnov, A.S., and Shchukin, E.D. (1992). Ostwald ripening theory: Applica- nanoemulsions prepared by high pressure homogenization. Food Technology
tions to fluorocarbon emulsion stability. Advances in Colloid and Interface and Biotechnology, 47():336–342.
Science, 38:69–97. Martini, É., Fattal, E., M. de Oliveira, and Teixeira, H. (2008). Effect of
Kawada, H., Kume, T., Matsunaga, T., Iwai, H., Sano, T., and Shibayama, M. cationic lipid composition on properties of oligonucleotide/emulsion com-
(2010). Structure and rheology of a self-standing nanoemulsion. Langmuir, plexes: Physico-chemical and release studies. International Journal of Phar-
26(4):2430–2437. maceutics, 352(1–2):280–286.
Kelmann, R., Kuminek, G., Teixeira, H., and Koester, L. (2007). Carbamazepine Mason, T.G., Wilking, J.N., Meleson, K., Chang, C.B., and Graves, S.M. (2006).
parenteral nanoemulsions prepared by spontaneous emulsification process. Nanoemulsions: Formation, structure, and physical properties. Journal of
International Journal of Pharmaceutics, 342(1–2):231–239. Physics-Condensed Matter, 18(41):R635–R666.
Kentish, S., Wooster, T., Ashokkumar, M., Balachandran, S., Mawson, R., and Maynard, A., Aitken, R., Butz, T., Colvin, V., Donaldson, K., Oberdörster, G.,
Simons, L. (2008). The use of ultrasonics for nanoemulsion preparation. Philbert, M., Ryan, J., Seaton, A., and Stone, V. (2006). Safe handling of
Innovative Food Science & Emerging Technologies, 9(2):170–175. nanotechnology. Nature, 444(7117):267–269.
Kentish, S., Wooster, T.J., Ashokkumar, A., Balachandran, S., Mawson, R., and McClements, D. (2005a). Food Emulsions: Principles, Practices, and Tech-
Simons, L. (2006). The Use of Ultrasonics for Nanoemulsion Preparation. 3rd niques. CRC Press, Boca Raton, FL
328 D. J. McCLEMENTS AND J. RAO

McClements, D.J. (2002a). Colloidal basis of emulsion color. Current Opinion Okamura, E., Kimura, T., Nakahara, M., Tanaka, M., Handa, T., and Saito,
in Colloid & Interface Science, 7(5–6):451–455. H. (2001). C-13 Nmr method for the determination of peptide and protein
McClements, D.J. (2002b). Theoretical prediction of emulsion color. Advances binding sites in lipid bilayers and emulsions. Journal of Physical Chemistry
in Colloid and Interface Science, 97(1–3):63–89. B, 105(50):12616–12621.
McClements, D.J. (2005b). Food Emulsions: Principles, Practice, and Tech- Ozaki, A., Muromachi, A., Sumi, M., Sakai, Y., Morishita, K., and Okamoto,
niques. CRC Press, Boca Raton, FL. T. (2010). Emulsification of Coenzyme Q(10) Using gum arabic increases
McClements, D.J. (2010). Design of nano-laminated coatings to control bioavailability in rats and human and improves food-processing suitability.
bioavailability of lipophilic food components. Journal of Food Science, Journal of Nutritional Science and Vitaminology, 56(1):41–47.
75(1):R30–R42. Palamakula, A., and Khan, M. (2004). Evaluation of cytotoxicity of oils used in
McClements, D.J., Decker, E.A., and Park, Y. (2007). Physicochemical and coenzyme Q10 self-emulsifying drug delivery systems (Sedds). International
structural aspects of lipid digestion. In: Understanding and Controlling the Journal of Pharmaceutics, 273(1–2):63–73.
Microstructure of Complex Foods, pp. 483–503. McClements, D.J., Ed., CRC Pena, A.A., and Miller, C.A. (2002). Transient behavior of polydisperse
Press, Boca Raton, FL. emulsions undergoing mass transfer. Industrial & Engineering Chemistry
McClements, D.J., Decker, E.A., and Park, Y. (2009). Controlling lipid bioavail- Research, 41(25):6284–6296.
ability through physicochemical and structural approaches. Critical Reviews Pey, C.M., Maestro, A., Sole, I., Gonzalez, C., Solans, C., and Gutierrez,
in Food Science and Nutrition, 49(1):48–67. J.M. (2006). Optimization of nano-emulsions prepared by low-energy
McClements, D.J., Decker, E.A., and Weiss, J. (2007). Emulsion-based deliv- emulsification methods at constant temperature using a factorial design
ery systems for lipophilioc bioactive components. Journal of Food Science, study. Colloids and Surfaces a-Physicochemical and Engineering Aspects,
72(8):R109-R124. 288(1–3):144–150.
McClements, D.J., and Dungan, S.R. (1993). Factors that affect the rate of oil Phan, T.T., Harwell, J.H., and Sabatini, D.A. (2010). Effects of triglyceride
exchange between oil-in-water emulsion droplets stabilized by a nonionic molecular structure on optimum formulation of surfactant-oil-water systems.
surfactant-droplet size, surfactant concentration, and ionic-strength. Journal Journal of Surfactants and Detergents, 13(2):189–194.
of Physical Chemistry, 97(28):7304–7308. Porter, C.J.H., Pouton, C.W., Cuine, J.F., and Charman, W.N. (2006). Enhancing
McClements, D.J., and Li, Y. (2010). Structured emulsion-based delivery sys- Intestinal Drug Solubilisation Using Lipid-Based Delivery Systems. Annual
tems: Controlling the digestion and release of lipophilic food components. Meeting of the American-Association-of-Pharmaceutical-Scientists, San
Advances in Colloid and Interface Science, 159(2):213–228. Antonio, TX.
Mei, L.Y., Choi, S.J., Alamed, J., Henson, L., Popplewell, M., D.J. McClements, Porter, C.J.H., Trevaskis, N.L., and Charman, W.N. (2007). Lipids and
and Decker, E.A. (2010). Citral stability in oil-in-water emulsions with lipid-based formulations: Optimizing the oral delivery of lipophilic drugs.
solid or liquid octadecane. Journal of Agricultural and Food Chemistry, Nature Reviews Drug Discovery, 6():231–248.
58(1):533–536. Pouton, C.W., and H, C.J.. Porter (2006). Formulation of Lipid-Based Delivery
Miller, C.A. (1988). Spontaneous emulsification produced by diffusion: A Systems for Oral Administration: Materials, Methods and Strategies. Annual
review. Colloids and Surfaces, 29(1):89–102. Meeting of the American-Association-of-Pharmaceutical-Scientists, San
Mira, I., Zambrano, N., Tyrode, E., Márquez, L., Peña, A., Pizzino, A., and Antonio, TX.
Salager, J. (2003). Emulsion Catastrophic inversion from abnormal to normal Qian, C., and McClements, D. (2010). Formation of nanoemulsions stabilized
morphology. 2. Effect of the stirring intensity on the dynamic inversion by model food-grade emulsifiers using high pressure homogenization:
frontier. Industrial Engineering and Chemical Research, 42(1):57–61. Factors affecting particle size. Food Hydrocolloids, In Press, DOI:
Muller, R.H., and Keck, C.M. (2004). Challenges and solutions for the 10.1016/j.foodhyd.2010.09.017.
delivery of biotech drugs: A review of drug nanocrystal technology and lipid Quemada, D., and Berli, C. (2002). Energy of interaction in colloids and its
nanoparticles. Journal of Biotechnology, 113(1–3):151–170. implications in rheological modeling. Advances in Colloid and Interface
Muller, R.H., Mader, K., and Gohla, S. (2000). Solid lipid nanoparticles (Sln) Science, 98(1):51–85.
for controlled drug delivery: A review of the state of the art. European Queste, S., Salager, J.L., Strey, R., and Aubry, J.M. (2007). The Eacn scale for
Journal of Pharmaceutics and Biopharmaceutics, 50(1):161–177. oil classification revisited thanks to fish diagrams. Journal of Colloid and
Müller, R.H., Radtke, M., and Wissing, S.A. (2002). Solid lipid nanoparticles Interface Science, 312(1):98–107.
(sln) and nanostructured lipid carriers (Nlc) in cosmetic and dermatolog- Quinn, A., Such, G.K., Quinn, J.F., and Caruso, F. (2008). Polyelectrolyte blend
ical preparations. Advanced Drug Delivery Review, 54 (Suppl 1):S131– multilayers: A versatile route to engineering interfaces and films. Advanced
155. Functional Materials, 18(1):17–26.
Mun, S., Decker, E.A., and McClements, D.J. (2005). Influence of droplet Quirantes, A., Plaza, R., and Delgado, A. (1997). Static light scattering study
characteristics on the formation of oil-in-water emulsions stabilized by of size parameters in core-shell colloidal systems. Journal of Colloid and
surfactant-chitosan layers. Langmuir, 21(14):6228–6234. Interface Science, 189(2):236–241.
Mun, S., Decker, E.A., Park, Y., Weiss, J., and McClements, D.J. (2006). Rao, J.J., and Mcclements, D.J. (2010). stabilization of phase inversion tem-
Influence of interfacial composition on in vitro digestibility of emulsified perature nanoemulsions by surfactant displacement. Journal of Agricultural
lipids: Potential mechanism for chitosan’s ability to inhibit fat digestion. and Food Chemistry, 58(11):7059–7066.
Food Biophysics, 1(1):21–29. Relkin, P., Jung, J.M., and Ollivon, M. (2009). Factors affecting vitamin
Mun, S.H., and McClements, D.J. (2006). Influence of interfacial characteristics degradation in oil-in-water nano-emulsions. Journal of Thermal Analysis
on Ostwald ripening in hydrocarbon oil-in-water emulsions. Langmuir, and Calorimetry, 98(1):13–18.
22(4):1551–1554. Ribeiro, H., Chu, B., Ichikawa, S., and Nakajima, M. (2008). Preparation of
Nakaya, K., Ushio, H., Matsukawa, S., Shimizu, M., and Ohshima, T. (2005). nanodispersions containing [beta]-carotene by solvent displacement method.
Effects of droplet size on the oxidative stability of oil-in-water emulsions. Food Hydrocolloids, 22(1):12–17.
Lipids, 40(5):501–507. Roger, E., Lagarce, F., Garcion, E., and Benoit, J.P. (2010). Biopharmaceutical
Nielsen, F.S., Petersen, K.B., and Mullertz, A. (2008). Bioavailability of parameters to consider in order to alter the fate of nanocarriers after oral
probucol from lipid and surfactant based formulations in minipigs: Influence delivery. Nanomedicine, 5(2):287–306.
of droplet size and dietary state. European Journal of Pharmaceutics and Rondón-Gonzaléz, M., Sadtler, V., Choplin, L., and Salager, J. (2006).
Biopharmaceutics, 69(2):553–562. Emulsion catastrophic inversion from abnormal to normal morphology.
Ochs, C.J., Such, G.K., Yan, Y., M.P. van Koeverden, and Caruso, F. (2010). 5. Effect of the water-to-oil ratio and surfactant concentration on the
Biodegradable click capsules with engineered drug-loaded multilayers. ACS inversion produced by continuous stirring. Ind. Eng. Chem. Res, 45(9):3074–
Nano, 4():1653–1663. 3080.
FOOD-GRADE NANOEMULSIONS 329

Sadtler, V., M. Rondon-Gonzalez, Acrement, A., Choplin, L., and Marie, E. Tadros, T., Izquierdo, R., Esquena, J., and Solans, C. (2004). Formation and
(2010). Peo-covered nanoparticles by emulsion inversion point (Eip) method. stability of nano-emulsions. Advances in Colloid and Interface Science,
Macromolecular Rapid Communications, 31(11):998–1002. 108–109:303–318.
Sadurn, N., Solans, C., Azemar, N., and M. Garca-Celma (2005). Studies Takegami, S., Kitamura, K., Kawada, H., Matsumoto, Y., Kitade, T., Ishida,
on the formation of o/w nano-emulsions, by low-energy emulsification H., and Nagata, C. (2008). Preparation and characterization of a new lipid
methods, suitable for pharmaceutical applications. European Journal of nano-emulsion containing two cosurfactants, sodium palmitate for droplet
Pharmaceutical Sciences, 26(5):438–445. size reduction and sucrose palmitate for stability enhancement. Chemical &
Sahay, G., Alakhova, D.Y., and Kabanov, A.V. (2010). Endocytosis of Pharmaceutical Bulletin, 56(8):1097–1102.
nanomedicines. Journal of Controlled Release, 145():182–195. Talegaonkar, S., Mustafa, G., Akhter, S., and Iqbal, Z.I. (2010). Design and
Sahay, G., Kim, J.O., Kabanov, A.V., and Bronich, T.K. (2010). The exploita- development of oral oil-in-water nanoemulsion formulation bearing atorvas-
tion of differential endocytic pathways in normal and tumor cells in the tatin: In vitro assessment. Journal of Dispersion Science and Technology,
selective targeting of nanoparticulate chemotherapeutic agents. Biomaterials, 31(5):690–701.
31(5):923–933. Tan, C.P., and Nakajima, M. (2005a). Beta-carotene nanodispersions:
Saito, H., Minamida, T., Arimoto, I., Handa, T., and Miyajima, K. (1996). Preparation, characterization and stability evaluation. Food Chemistry,
Physical states of surface and core lipids in lipid emulsions and apolipopro- 92(4):661–671.
tein binding to the emulsion surface. Journal of Biological Chemistry, Tan, C.P., and Nakajima, M. (2005b). Effect of polyglycerol esters of fatty acids
271(26):15515–15520. on physicochemical properties and stability of beta-carotene nanodispersions
Salager, J., Antón, R., Sabatini, D., Harwell, J., Acosta, E., and Tolosa, L. prepared by emulsification/evaporation method. Journal of the Science of
(2005). Enhancing solubilization in microemulsions—state of the art and Food and Agriculture, 85(1):121–126.
current trends. Journal of Surfactants and Detergents, 8(1):3–21. Tanaka, M., Saito, H., Arimoto, I., Nakano, M., and Handa, T. (2003).
Salager, J.L., Forgiarini, A., Marquez, L., Pena, A., Pizzino, A., Rodriguez, Evidence for interpenetration of core triglycerides into surface phospholipid
M.P., and M. Rondo-Gonzalez (2004). Using emulsion inversion in in- monolayers in lipid emulsions. Langmuir, 19(13):5192–5196.
dustrial processes. Advances in Colloid and Interface Science, 108:259– Teixeira, P.C., Leite, G.M., Domingues, R.J., Silva, J., Gibbs, P.A., and Ferreira,
272. J.P. (2007). Antimicrobial effects of a microemulsion and a nanoemulsion
Sandra, S., Decker, E.A., and McClements, D.J. (2008). Effect of interfacial pro- on enteric and other pathogens and biofilms. International Journal of Food
tein cross-linking on the in vitro digestibility of emulsified corn oil by pancre- Microbiology, 118(1):15–19.
atic lipase. Journal of Agricultural and Food Chemistry, 56(16):7488–7494. Teo, B.S.X., Basri, M., M.R.S. Zakaria., Salleh, A.B., R.N.Z.R.A. Rahman,
Schubert, H., Ax, K., and Behrend, O. (2003). Product engineering of dispersed and A, M.B.. Rahman (2010). A potential tocopherol acetate loaded palm
systems. Trends in Food Science & Technology, 14(1):9–16. oil esters-in-water nanoemulsions for nanocosmeceuticals. Journal of
Schubert, H., and Engel, R. (2004). Product and formulation engineer- Nanobiotechnology, 8(4): 1–12.
ing of emulsions. Chemical Engineering Research & Design, 82(A9): Thakur, R.K., Villette, C., Aubry, J.M., and Delaplace, G. (2008). Dynamic
1137–1143. emulsification and catastrophic phase inversion of lecithin-based emul-
Sekhon, B. (2010). Food nanotechnology – an overview. Nanotechnology, sions. Colloids and Surfaces A-Physicochemical and Engineering Aspects,
Science and Applications, 3:1–15. 315(1–3):285–293.
Shafiq-un-Nabi, S., Shakeel, F., Talegaonkar, S., Ali, J., Baboota, S., Ahuja, Trotta, M., Cavalli, R., Ugazio, E., and Gasco, M. (1996). Phase Behaviour of
A., Khar, R., and Ali, M. (2007). formulation development and optimization Microemulsion systems containing lecithin and lysolecithin as surfactants.
using nanoemulsion technique: A technical note. AAPS PharmSciTech, 8(2): International Journal of Pharmaceutics, 143(1):67–73.
12–17. Turnbull, D., and Cormia, R.L. (1961). Kinetics of crystal nucleation in some
Shen, Z.P., Udabage, P., Burgar, I., and Augustin, M.A. (2005). Characteriza- normal alkane liquids. Journal of Chemical Physics, 34(3):820–828.
tion of fish oil-in-water emulsions using light-scattering, nuclear magnetic Unfried, K., Albrecht, C., L.O. Klotz, A. Von Mikecz, S. Grether-Beck, and
resonance, and gas chromatography-headspace analyses. Journal of the F, R.P.. Schins (2007). Cellular responses to nanoparticles: Target structures
American Oil Chemists Society, 82(11):797–802. and mechanisms. Nanotoxicology, 1(1):52–71.
Silva, H., Cerqueira, M., Souza, B., Ribeiro, C., Avides, M., Quintas, M., van Aken, G.A. (2010a). Modelling texture perception by soft epithelial
Coimbra, J., M. Carneiro-da-Cunha, and Vicente, A. (2011). Nanoemulsions surfaces. Soft Matter, 6(5):826–834.
of [beta]-carotene using a high-energy emulsification-evaporation technique. van Aken, G.A. (2010b). Relating food emulsion structure and composition
Journal of Food Engineering. 102(2): 130–135. to the way it is processed in the gastrointestinal tract and physiologi-
Singh, H., Ye, A.Q., and Horne, D. (2009). Structuring food emulsions in the cal responses: What are the opportunities? Food Biophysics, 5(4):258–
gastrointestinal tract to modify lipid digestion. Progress in Lipid Research, 283.
48(2):92–100. Velikov, K.P., and Pelan, E. (2008). Colloidal delivery systems for micronutri-
Sol, I., Maestro, A., Pey, C., González, C., Solans, C., and Guti, J.érrez (2006). ents and nutraceuticals. Soft Matter, 4(10):1964–1980.
Nano-emulsions preparation by low-energy methods in an ionic surfactant Versantvoort, C.H.M., E. Van de Kamp, and M, C.J.. Rompelberg (2004).
system. Colloids and Surfaces A: Physicochemical and Engineering Aspects, Development and applicability of an in vitro digestion model in assessing
288(1–3):138–143. the bioaccessibility of contaminants from food. Bilthoven, Inspectorate of
Solè, I., Maestro, A., González, C., Solans, C., and J. Gutiérrez (2006). Health Inspection, pp. 1–87.
Optimization of nano-emulsion preparation by low-energy methods in an Vishwanathan, R., Wilson, T., and Nicolosi, R. (2009). Bioavailability of a
ionic surfactant system. Langmuir, 22(20):8326–8332. nanoemulsion of lutein is greater than a lutein supplement. Nano Biomedicine
Sonneville-Aubrun, O., Babayan, D., Bordeaux, D., Lindner, P., Rata, G., and and Engineering, 1(1):53.
Cabane, B. (2009). Phase transition pathways for the production of 100 nm oil- Walstra, P. (1993). Principles of emulsion formation. Chemical Engineering
in-water emulsions. Physical Chemistry Chemical Physics, 11(1):101–110. Science, 48(2): 333–349.
Sonneville-Aubrun, O., Simonnet, J.T., and F. L’Alloret (2004). Nanoemul- Walstra, P. (2003). Physical Chemistry of Foods. Marcel Decker, New York.
sions: A new vehicle for skincare products. Advances in Colloid and Wang, X.Y., Jiang, Y., Wang, Y.W., Huang, M.T., Ho, C.T., and Huang, Q.R.
Interface Science, 108:145–149. (2008). Enhancing anti-inflammation activity of curcumin through O/W
Souto, E.B., P. Martins-Lopes, Lopes, C.M., Gaivao, I., Silva, A.M., and nanoemulsions. Food Chemistry, 108(2):419–424.
H. Guedes-Pinto (2009). A note on regulatory concerns and toxicity Weiss, J., Canceliere, C., and McClements, D.J. (2000). Mass transport phe-
assessment in lipid-based delivery systems (Lds). Journal of Biomedical nomena in oil-in-water emulsions containing surfactant micelles: Ostwald
Nanotechnology, 5(4):317–322. ripening. Langmuir, 16(17):6833–6838.
330 D. J. McCLEMENTS AND J. RAO

Weiss, J., Decker, E., A., D.J. McClements, Kristbergsson, K., Helga- Wooster, T., Golding, M., and Sanguansri, P. (2008). Impact of oil type
son, T., and Awad, T. (2007). Solid Lipid Nanoparticles as Delivery on nanoemulsion formation and Ostwald ripening stability. Langmuir,
Systems for Bioactive Food Components. 2nd International Sympo- 24(22):12758–12765.
sium on Delivery of Functionality in Complex Food Systems, Amherst, Wooster, T.J., Golding, M., and Sanguansri, P. (2008a). Impact of oil type
MA. on nanoemulsion formation and Ostwald ripening stability. Langmuir,
Weiss, J., Decker, E.A., D.J. McClements, Kristbergsson, K., Helgason, T., and 24(22):12758–12765.
Awad, T. (2008). Solid lipid nanoparticles as delivery systems for bioactive Wulff-Perez, M., M.J. Galvez-Ruiz, J. de Vicente, and A. Martin-Rodriguez
food components. Food Biophysics, 3(2):146–154. (2010). Delaying lipid digestion through steric surfactant Pluronic F68: A
Weiss, J., and McClements, D.J. (2000). Influence of Ostwald ripening on novel in vitro approach. Food Research International, 43(6):1629–1633.
rheology of oil-in-water emulsions containing electrostatically stabilized Wulff-Perez, M., A. Torcello-Gomez, M.J. Galvez-Ruiz, and A. Martin-
droplets. Langmuir, 16(5):2145–2150. Rodriguez (2009). Stability of emulsions for parenteral feeding: Preparation
Weiss, J., and McClements, D.J. (2001). Color changes in hydrocarbon and characterization of O/W nanoemulsions with natural oils and pluronic
oil-in-water emulsions caused by Ostwald ripening. Journal of Agricultural F68 as surfactant. Food Hydrocolloids, 23(4):1096–1102.
and Food Chemistry, 49(9):4372–4377. Yaghmur, A., Aserin, A., and Garti, N. (2002). Phase behavior of microemul-
Westesen, K. (2000). Novel lipid-based colloidal dispersions as potential drug sions based on food-grade nonionic surfactants: Effect of polyols and
administration systems – Expectations and reality. Colloid and Polymer short-chain alcohols 1. Colloids and Surfaces A: Physicochemical and
Science, 278(7):608–618. Engineering Aspects, 209(1):71–81.
Wilking, J.N., and Mason, T.G. (2007). Irreversible shear-induced vitrification Yin, L., Chu, B., Kobayashi, I., and Nakajima, M. (2009). Performance
of droplets into elastic nanoemulsions by extreme rupturing. Physical Review of selected emulsifiers and their combinations in the preparation of
E, 75(4):5. [beta]-carotene nanodispersions. Food Hydrocolloids, 23(6):1617–1622.
Wissing, S.A., Kayser, O., and Muller, R.H. (2004). Solid lipid nanoparticles Yuan, H., Chen, J., Du, Y.Z., Hu, F.Q., Zeng, S., and Zhao, H.L. (2007).
for parenteral drug delivery. Advanced Drug Delivery Reviews, 56(9):1257– Studies on oral absorption of stearic acid sln by a novel fluorometric method.
1272. Colloids and Surfaces B-Biointerfaces, 58(2):157–164.
Witthayapanyanon, A., Acosta, E., Harwell, J., and Sabatini, D. (2006). For- Yuan, Y., Gao, Y., Zhao, J., and Mao, L. (2008a). Characterization and stability
mulation of ultralow interfacial tension systems using extended surfactants. evaluation of [beta]-carotene nanoemulsions prepared by high pressure
Journal of Surfactants and Detergents, 9(4):331–339. homogenization under various emulsifying conditions. Food Research
Witthayapanyanon, A., Harwell, J., and Sabatini, D. (2008). Hydrophilic- International, 41(1):61–68.
Lipophilic Deviation (Hld) Method for characterizing conventional and Yuan, Y., Gao, Y.X., Mao, L., and Zhao, J. (2008b). Optimisation of conditions
extended surfactants. Journal of Colloid and Interface Science, 325(1):259– for the preparation of beta-carotene nanoemulsions using response surface
266. methodology. Food Chemistry, 107():1300–1306.

You might also like