EL 2014 Antimicrobial Activity of Emulsions and Nanoemulsions

You might also like

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

Food Hydrocolloids xxx (2014) 1e10

Contents lists available at ScienceDirect

Food Hydrocolloids
journal homepage: www.elsevier.com/locate/foodhyd

Physicochemical characterization and antimicrobial activity of food-


grade emulsions and nanoemulsions incorporating essential oils
Laura Salvia-Trujillo, Alejandra Rojas-Graü, Robert Soliva-Fortuny, Olga Martín-Belloso*
Department of Food Technology, University of Lleida e Agrotecnio Center, Av. Alcalde Rovira Roure 191, 25198, Lleida, Spain

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

Article history: Coarse emulsions containing essential oils (lemongrass, clove, tea tree, thyme, geranium, marjoram,
Received 21 November 2013 palmarosa, rosewood, sage or mint) and stabilized with tween 80 and sodium alginate were prepared by
Accepted 13 July 2014 high shear homogenization. Nanoemulsions were obtained by microfluidization of coarse emulsions. In
Available online xxx
general, the average droplet size of coarse emulsions was dramatically reduced after microfluidization
down to a few nanometers, with the exception of palmarosa and rosewood oil emulsions, which were
Keywords:
already in the nano-range before being treated. The z-potential of nanoemulsions exhibited values more
Antimicrobial activity
negative than 30 mV, indicating a strong electrostatic repulsion of the dispersed oil droplets in the
Escherichia coli
Nanoemulsion
aqueous phase. The viscosity of nanoemulsions significantly decreased after microfluidization, with at
Essential oils least a 30% drop in their initial values. The whiteness index of nanoemulsions diminished after being
Microfluidization treated. In fact, nanoemulsions containing tea tree, geranium or marjoram essential oils became
Droplet size completely transparent after microfluidization. Lemongrass, clove, thyme or palmarosa-loaded nano-
emulsions were those with a higher in vitro bactericidal action against Escherichia coli, as they achieved
4.1, 3.6, 2.8 or 3.9 log-reductions after 30 min of contact time. In addition, a faster and enhanced inac-
tivation kinetic was observed in the case of nanoemulsions containing lemongrass or clove essential oils
in comparison with their respective coarse emulsions. Thus, the present work evidences the promising
advantages of using nanoemulsions as delivery systems of flavoring and preservative agents in the food
industry.
© 2014 Elsevier Ltd. All rights reserved.

1. Introduction the incorporation of antimicrobial essential oils to foods still pre-


sents several drawbacks due to their poor water solubility as well as
Essential oils contain a complex mixture of non-volatile and to toxicological and economical considerations (Burt, 2004;
volatile compounds produced by aromatic plants as secondary Sanchez-Gonz alez, Vargas, Gonza lez-Martínez, Chiralt, & Cha fer,
metabolites (Bakkali, Averbeck, Averbeck, & Idaomar, 2008; Burt, 2011). In addition, their strong flavor makes incorporation at high
2004). They are used in a wide variety of applications in food, doses difficult in certain types of food products due to possible
pharmaceutical and cosmetics industries due to their flavoring, objectionable sensory characteristics. Hence, there is a need to
antioxidant and antimicrobial properties (Adorjan & Buchbauer, investigate new delivery systems to encapsulate and release
2010; Brud, 2010; Tajkarimi, Ibrahim, & Cliver, 2010). In partic- essential oils in food products.
ular, the antimicrobial action of essential oils has been attributed to In the food sector, the incursion of nanotechnological advances
their phenolic compounds and their interaction with microbial cell is still discreet but it is gaining more and more interest by both the
membranes. They are known to penetrate through the microbial scientific and industrial community (Rashidi & Khosravi-Darani,
membrane and cause the leakage of ions and cytoplasmatic content 2011). Recently, emulsions with small droplet size, typically from
thus leading to cellular breakdown (Burt, 2004). The interest of 10 to 100 nm, also called nanoemulsions, are being investigated as
incorporating essential oils in foods as preservatives is related to lipophilic drug delivery systems in food, cosmetic and pharma-
their recognition as safe natural compounds, being a potential ceutical products (Bernardi et al., 2011; He et al., 2011). Due to their
alternative to produce foods free of synthetic additives. However, intrinsic properties, they may present several advantages for
encapsulating functional lipophilic compounds over conventional
emulsions. On the one hand, their reduced droplet size might not
* Corresponding author.
only enhance the transport of active molecules through biological
E-mail address: omartin@tecal.udl.es (O. Martín-Belloso).

http://dx.doi.org/10.1016/j.foodhyd.2014.07.012
0268-005X/© 2014 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Salvia-Trujillo, L., et al., Physicochemical characterization and antimicrobial activity of food-grade emulsions
and nanoemulsions incorporating essential oils, Food Hydrocolloids (2014), http://dx.doi.org/10.1016/j.foodhyd.2014.07.012
2 L. Salvia-Trujillo et al. / Food Hydrocolloids xxx (2014) 1e10

membranes but also increase the surface area/volume ratio, thus according to the information given by the supplier. Tween 80 (1% v/
leading to improved functionality. On the other hand, nano- v) (Scharlau, Spain) was added as surfactant with a laboratory T25
emulsions are kinetically stable and transparent colloidal disper- digital Ultra-Turrax mixer (IKA, Staufen, Germany) working at
sions, being suitable for a wide range of practical applications 3400 rpm for 2 min. All samples were prepared using ultra pure
(Solans, Izquierdo, Nolla, Azemar, & Garcia-Celma, 2005). water obtained from a Mili-Q filtration system.
To produce nanoemulsions with droplet sizes in the nano-range
a high shear rate is required in order to generate a high interface 2.2. Nanoemulsion formation
area (Delmas et al., 2011). Among the few methods that can pro-
duce nanoemulsions, microfluidization has been shown to generate A microfluidization system (M110P, Microfluidics, Massachu-
extremely fine emulsions with droplet sizes between 60 and setts, USA) was used to obtain nanoemulsions from the coarse
600 nm (Hatanaka et al., 2010; Hatanaka, Kimura, Lai-Fu, Onoue, & emulsions. Coarse emulsions were passed through the system 3
Yamada, 2008; Jafari, He, & Bhandari, 2007a, 2007b; Qian & times at a constant pressure of 150 MPa. At the outlet of the
McClements, 2011; Rao & McClements, 2011; Wooster, Golding, & interaction chamber the product was refrigerated through an
Sanguansri, 2008). The microfluidizer device uses a pump to force a external cooling coil immersed in an ice-water bath so that the
coarse emulsion pre-mix to an interaction chamber under high temperature of the product was always kept below 20  C.
pressure (Jafari et al., 2007a; McClements, 2011). In the interaction
chamber the coarse emulsion is split in two streams that further 2.3. Emulsion and nanoemulsion characterization
impinge on each other at high velocity, providing an exceptionally
fine emulsion (Mahdi Jafari, He, & Bhandari, 2006; McClements, 2.3.1. Particle size and z-potential
2011). The oil droplet size was measured by dynamic light scattering
The use of nanoemulsions as potential delivery systems of (DLS) with a Zetasizer NanoZS laser diffractometer (Malvern In-
lipophilic food ingredients is arising with promising expectations. struments Ltd, Worcestershire, UK) working at 633 nm at 25  C and
However, there is a lack of scientific evidence about the enhanced equipped with a backscatter detector (173 ), which is appropriate
functionality of nanoemulsions in comparison with similar non- to measure sub-micron particles (Brar & Verma, 2011). DLS mea-
nano-formulated delivery systems (Bouwmeester et al., 2009). sures the Brownian motion of nano-sized droplets and relates this
Therefore, the purpose of the present research work was to char- movement to an equivalent hydrodynamic diameter (nm). Average
acterize microfluidized nanoemulsions incorporating essential oils droplet size, size distribution curves and polydispersity index were
in terms of droplet size, size distribution, z-potential, viscosity and used to characterize oil droplets dispersion in nanoemulsions. The
color in comparison with conventional emulsions. Moreover, the refractive index of essential oils and their absorbance at 633 nm,
in vitro antimicrobial activity against Escherichia coli of nano- were measured with a refractometer and a spectrophotometer,
emulsions was assessed and compared with that exhibited by respectively (Table 1).
conventional emulsions. The electrophoretic mobility of oil droplets, also reported as z-
potential, was measured by phase-analysis light scattering (PALS)
2. Material and methods with a Zetasizer NanoZS laser diffractometer (Malvern Instruments
Ltd, Worcestershire, UK). It determines the surface electrical charge
2.1. Primary emulsion formation of the droplets dispersed in the continuous phase.

Sodium alginate (FMC Biopolymers, UK) was dissolved in hot 2.3.2. Viscosity
water at 70  C (1% w/v) under continuous stirring until complete Viscosity of emulsions and nanoemulsions was measured from
homogenization. Coarse emulsions were prepared by mixing the approximately 10 mL of sample using a SV-10 vibro-viscometer
resulting sodium alginate aqueous solution with essential oils (1% (A&D Company, Tokyo, Japan) vibrating at 30 Hz and constant
v/v). Sodium alginate was included in the aqueous phase as the amplitude. The viscosity of the sodium alginate solution (30 mPa s)
purpose of further investigations was to apply such nanoemulsions was considered the viscosity of the dispersant material with
as antimicrobial edible coatings in fresh-cut fruit. Lemongrass respect to the dynamic light scattering measurements.
(Cymbopogon citratus) essential oil was purchased at Laboratoris
Dicana (Spain) and clove (Eugenia caryophyllata), tea tree (Mela- 2.3.3. Whiteness index
leuca alternifolia), thyme (Thymus vulgaris), geranium (Pelargonium The color of emulsions and nanoemulsions was measured with a
graveolens), marjoram (Origanum majorana), palmarosa (Cymbo- Minolta CR-400 colorimeter (Konica Minolta Sensing, Inc., Osaka,
pogon martinii), rosewood (Aniba rosaedora), sage (Salvia officinalis) Japan) at room temperature set up for illuminant D65 and 10
and mint (Mentha spicata) were purchased from Dietetica Intersa observer angle and calibrated with a standard white plate. CIE L*, a*
(Spain). The composition of each type of oil is detailed in Table 1, and b* values were determined and the Whiteness Index (WI) was

Table 1
Essential oil composition, oil refractive index and absorbance (l ¼ 633 nm) of essential oils.

Essential oil Plant Composition Refractive index Abs633

Lemongrass Cymbopogon citratus Geranial, neral, geraniol 1.453 0.022


Clove Eugenia caryophyllata Eugenol, eugenil acetate 1.499 0.023
Tea Tree Melaleuca alternifolia 4-terpineol, terpinene 1.465 0.001
Thyme Thymus vulgaris Timol, p-cimene 1.471 0.094
Geranium Pelargonium graveolens Citronelol, geraniol 1.462 0.157
Marjoram Origanum majorana 4-terpineol, cis tuyan-4-ol 1.472 0.002
Palmarosa Cymbopogon martinii Geraniol, geranil acetate, linalol 1.461 0.003
Rosewood Aniba rosaedora a-terpineol, linalol 1.468 0.002
Sage Salvia officinalis Canfor, 1,8-cineol, canfene, sabinil acetate 1.479 0.071
Mint Mentha spicata Mentol 1.477 0.002

Please cite this article in press as: Salvia-Trujillo, L., et al., Physicochemical characterization and antimicrobial activity of food-grade emulsions
and nanoemulsions incorporating essential oils, Food Hydrocolloids (2014), http://dx.doi.org/10.1016/j.foodhyd.2014.07.012
L. Salvia-Trujillo et al. / Food Hydrocolloids xxx (2014) 1e10 3

calculated with equation (1) (Vargas, Ch


afer, Albors, Chiralt, & and nanoemulsions containing different essential oils at a 5% sig-
Gonzalez-Martínez, 2008): nificance level.
  0:5
WI ¼ 100  ð100  LÞ2 þ a2 þ b2 (1) 3. Results

3.1. Particle size and particle size distribution


2.3.4. Antimicrobial activity assay
The antimicrobial activity of essential oil-alginate emulsions The results of the average droplet size and polydispersity index
and nanoemulsions was assessed evaluating the in vitro inhibition of essential oil-loaded emulsions and nanoemulsions are shown in
of E. coli, following the method described previously by other au- Table 2, and their droplet size distribution is presented in Fig. 1. The
thors with some modifications (Ferreira et al., 2010). E. coli 1.107 method used to prepare the mixtures significantly influenced the
(Laboratoire de Re pression des Fraudes, Montpellier, France) was emulsion droplet size. The further processing of blends by micro-
cultured in tryptone soy broth (Bioakar Diagnostic; Beauvais, fluidization after a previous high shear homogenization rendered
France) and incubated at 37  C with continuous agitation at nanoemulsions with a reduced droplet size. Moreover, strong dif-
120 rpm for 11 h to obtain cells in stationary growth phase. The ferences among emulsion droplet size and size distribution were
final concentration reached in the culture was 108e109 colony observed depending on the essential oil used in the formulation of
forming units/millilitre (CFU/mL). A 0.5 mL-aliquot of overnight nanoemulsions. Namely, microfluidized lemongrass oil-loaded
bacterial culture was mixed with 0.5 mL of the essential oil-alginate nanoemulsions exhibited a dramatic reduction of their average
emulsion or nanoemulsion and 4.5 mL of sterile Mili-Q water. Ali- droplet size, down to 5.5 ± 0.3 nm, in comparison with the values of
quots of 0.1 mL were taken after 5, 15 and 30 min and were serially the coarse emulsion, which averaged 1120 ± 230 nm. Oppositely,
diluted and spread on McConkey agar (Biokar Diagnostics, Beau- emulsions incorporating palmarosa and rosewood oils exhibited
vais, France) plates to determine the inactivation kinetics. A blank average droplet sizes already in the nano-scale range (33 ± 7 or
determination was performed with the same method, replacing the 13.8 ± 0.5 nm, respectively) before microfluidization and with a
emulsion or nanoemulsion by sterile Mili-Q water. Colony counts narrow size distribution. Therefore, in both cases the micro-
were determined after incubation of agar plates at 37  C for 24 h. fluidization treatment had a slight effect on the average droplet
size, which was confirmed by size distribution data (Fig. 1).
2.4. Kinetic modeling Regarding the rest of essential oils (clove, tea tree, thyme, gera-
nium, marjoram, sage or mint), the average droplet size of coarse
The inactivation kinetics of essential oil-alginate emulsions or emulsions was also in the nano-scale range, with values between
nanoemulsions against E. coli was adjusted to a Weibull distribu- 13.2 ± 0.9 and 87 ± 4 nm. Nonetheless, a multimodal size distri-
tion model, by Eq. (2) (Peleg, 2006): bution was observed with major intensity peaks positioned at
 a 1000 nm and with minor peaks at the nano-range. Those minor
t peaks might be surfactant micelles that were not adsorbed in the
ln SðtÞ ¼  (2)
b oil droplets interface, which are typically around 10 nm
(Heydenreich, Westmeier, Pedersen, Poulsen, & Kristensen, 2003;
where S is the survival fraction (N/N0) of E. coli, t is the reaction time Menard et al., 2011).
(min), b is the scale factor, and a is the shape factor. The scale factor In fact, the particle size distribution graphs of microfluidized
represents the contact time (min) necessary to inactivate the first emulsions (Fig. 1) show a significant decrease in the intensity peaks
log-cycle of the microbial population, whereas the shape factor is of particles above 100 nm. Consequently, the intensity peaks of
related to the trend of the inactivation curve. A shape factor below 1 smaller particles increased, indicating a clear particle size
indicates a concavity upwards while a value greater than 1 means a
concavity downwards.
The model was fitted to experimental data by non-linear Table 2
Average droplet size (z-average) (nm) and polydispersity index (PDI) of essential oil-
regression using the Statgraphics Plus v.5.1 Windows package.
alginate emulsions or nanoemulsions produced by high shear homogenization
The adjusted regression coefficients (R2 adj) and accuracy factor (Af, (3400 rpm, 2 min) or microfluidization (150 MPa, 3 cycles).
Eq. (3)) (Ross, 1996) of the model were calculated.
Essential oil Homogenization system z-average (nm) PDI
  
P predicated  Lemongrass HSH 1121.43 ± 228.93 0.49 ± 0.08
 log
observed  MF 5.49 ± 0.28 0.51 ± 0.04
Af ¼ 10 n (3) Clove HSH 23.23 ± 10.51 0.94 ± 0.09
MF 20.88 ± 5.10 0.82 ± 0.07
Tea Tree HSH 16.20 ± 5.24 1.91 ± 0.70
where log (predicted/observed) is the logarithmic relation between
MF 2.65 ± 0.68 0.10 ± 0.01
the predicted and the experimental values, and n is the number of Thyme HSH 13.16 ± 0.92 0.23 ± 0.03
observations. MF 12.86 ± 0.25 0.36 ± 0.01
The higher the values of R2 adj and the nearer the values of Af to Geranium HSH 29.24 ± 9.73 0.38 ± 0.18
1, the better the fitting of the model to experimental data. MF 2.88 ± 0.16 0.28 ± 0.09
Marjoram HSH 35.14 ± 8.26 0.21 ± 0.18
MF 7.71 ± 0.64 0.10 ± 0.01
2.5. Statistics Palmarosa HSH 32.91 ± 6.92 1.02 ± 0.01
MF 13.82 ± 0.31 0.71 ± 0.08
Rosewood HSH 13.82 ± 0.48 1.02 ± 0.32
All the experiments were assayed in duplicate, and two replicate
MF 12.26 ± 0.09 1.08 ± 0.11
analyses of each sample were carried out in order to calculate a Sage HSH 86.49 ± 4.31 0.27 ± 0.09
mean value. Statgraphics Plus v.5.1 Windows package (Statistical MF 8.56 ± 1.29 0.34 ± 0.04
Graphics Co., Rockville, MD) was used to perform the analysis of Mint HSH 13.95 ± 1.97 0.78 ± 0.01
variance. The least significant difference (LSD) test was run to MF 2.22 ± 0.06 0.12 ± 0.01

determine significant differences (p  0.05) between emulsions Values are expressed as mean ± standard deviation.

Please cite this article in press as: Salvia-Trujillo, L., et al., Physicochemical characterization and antimicrobial activity of food-grade emulsions
and nanoemulsions incorporating essential oils, Food Hydrocolloids (2014), http://dx.doi.org/10.1016/j.foodhyd.2014.07.012
4 L. Salvia-Trujillo et al. / Food Hydrocolloids xxx (2014) 1e10

Fig. 1. Droplet size distribution of essential oil-alginate emulsions or nanoemulsions produced by high shear homogenization (3400 rpm, 2 min) (continuous line) or micro-
fluidization (150 MPa, 3 cycles) (dotted line).

reduction. The average droplet size (z-average) of emulsions cor- impact forces as well as cavitation phenomena within the product
responds to the mean hydrodynamic diameter and is calculated that reduce the emulsion droplet size (Maa & Hsu, 1999). However,
according to the ISO-13321 (1996), but it may lead to misinter- the final droplet size of nanoemulsions is a complex interplay be-
pretation in the case of non-monodispersed emulsions. In general, tween the processing conditions and the adsorption of surfactant
from the results described above it can be observed a clear droplet onto oil droplets (Wooster et al., 2008). On the one hand, the ability
size disruption caused by microfluidization treatment, leading to of flavor oil molecules to be incorporated into surfactant micelles
nanoemulsions with small droplet sizes. During the micro- depends on their molecular characteristics, such as molecular
fluidization treatment, the process stream is pressurized and weight, polarity and conformation (Djekic & Primorac, 2008). It has
passed through an interaction chamber thus creating shear and been reported that essential oils with higher concentration of polar

Please cite this article in press as: Salvia-Trujillo, L., et al., Physicochemical characterization and antimicrobial activity of food-grade emulsions
and nanoemulsions incorporating essential oils, Food Hydrocolloids (2014), http://dx.doi.org/10.1016/j.foodhyd.2014.07.012
L. Salvia-Trujillo et al. / Food Hydrocolloids xxx (2014) 1e10 5

compounds might reduce the interfacial tension and facilitate nanoemulsion system (Heurtault, Saulnier, Pech, Proust, & Benoit,
droplet disruption during high pressure homogenization (Ziani, 2003). On one hand, coarse emulsions formed by high shear ho-
Fang, & McClements, 2012). In this sense, low molecular weight mogenization exhibited a weak surface charge with values less
or high polarity compounds tend to be solubilized in the aqueous negative than 30 mV. In fact, the coarse emulsion incorporating
phase rather than in the micelle interior. On the other hand, the mint essential oil exhibited the weaker z-potential (10.0 mV). In
emulsifier type and concentration are important in determining the contrast, clove, tea tree or marjoram-loaded emulsions treated by
emulsion droplet size, since they should adsorb to oil droplet sur- high shear homogenization exhibited a strong electrical charge
face to prevent re-coalescence (Jafari, Assadpoor, He, & Bhandari, of 52.6, 30.8 and 37.4 mV, respectively. On the other hand,
2008). Moreover, viscosity and concentration of the dispersed nanoemulsions obtained by microfluidization showed a significant
phase are determining factors when considering the formation of decrease in their z-potential regardless of the oil type used in the
nano-sized emulsions (Jafari et al., 2008; McClements, 2005; Qian formulation. In fact, most of the nanoemulsions had z-potential
& McClements, 2011; Seekkuarachchi, Tanaka, & Kumazawa, values more negative than 30 mV, with the exception of
2006; Wooster et al., 2008). palmarosa-loaded nanoemulsions (20.9 mV). According to
It has been observed that the higher the viscosity of the oil s, Vargas, and Chiralt (2012), differences observed in
Bonilla, Atare
phase, the higher the emulsion droplet size (Qian & McClements, the z-potential of emulsions and nanoemulsions formulated with
2011). In fact, as the oil viscosity increases, the disruption forces different essential oils might be attributed to differences in the
needed to deform the oil droplets during homogenization are dissociation degree and number of ionizable compounds of oils.
higher (Håkansson, Tra €gårdh, & Bergenståhl, 2009; McClements, Moreover, it is generally considered that the competitive adsorp-
2005). Namely, Wooster et al. (2008) observed droplet sizes of tion of the different emulsifiers is based on principle that the more
around 120 nm in nanoemulsions formulated with high viscosity surface-active material replaces the less active compound
oils, whereas nanoemulsions containing low viscosity oils pre- (Hasenhuettl & Hartel, 2008; Kralova & Sjo € blom, 2009), but this
sented droplet sizes of around 80 nm. Usually, plant essential oils might depend on the affinity with the type of oil. Therefore, the
are constituted of low molecular weight compounds and present differences in the z-potential observed among essential oils might
low viscosity values (Sell, 2010). In agreement to our results, Rao be due to differences between the adsorption of the surface-active
and McClements (2012) and Ziani, Chang, McLandsborough, and compounds in the oilewater interface.
McClements (2011) have used microfluidization approaches for Tween 80 is a non-ionic low-mass surfactant that rapidly coats
the generation of lemon oil and thyme oil-loaded nanoemulsions, the newly created droplets during emulsification (Kralova &
respectively. In addition, Donsì, Annunziata, Sessa, and Ferrari Sjo€ blom, 2009). It is preferably adsorbed on the oil surface in the
(2011) obtained D-limonene nanoemulsions stabilized with tween co-presence of polymeric stabilizers (Nambam & Philip, 2012),
20 and glycerol monooleate through high pressure homogenization giving a neutral (Kralova & Sjo €blom, 2009) or slightly negative
(10 cycles at 300 MPa) with an average droplet size of electrical charge on oilewater interface (Hsu & Nacu, 2003). Thus,
154.6 ± 1.4 nm. Differences in droplet size observed in the present the negative charge of oil droplets observed in the present work
work between various essential oil-loaded nanoemulsions might might be due to the anionic sodium alginate molecules dispersed in
be attributed to their molecular structure, concentration of volatile the aqueous phase. Even though alginates, as well as most food
compounds, viscosity, interfacial tension or surfactant affinity to hydrocolloids, do not possess hydrophobic moieties (Dickinson,
each type of flavor oil. 2003, 2009), it has recently been observed that they can present
surface activity under certain conditions (Garti & Leser, 2001; Yang,
Jiang, He, & Xia, 2012). The mechanical stress during micro-
3.2. z-potential
fluidization can lead to break up sodium alginate (see viscosity
discussion 3.3) thus increasing the number of molecules to be
The surface charge of oil droplets in emulsions and nano-
potentially adsorbed on the oilewater interface. This fact could
emulsions incorporating several essential oils is shown in Fig. 2.
explain the z-potential strengthening observed in microfluidized
Particles with z-potential more positive than þ30 mV or more
nanoemulsions.
negative than 30 mV are usually considered to be stable, since
electrical charge of droplets is strong enough to assume that
3.3. Viscosity
repulsive forces between droplets are predominant in the
The viscosity of emulsions or nanoemulsions incorporating
different essential oils before and after microfluidization at
150 MPa for 3 cycles is presented in Fig. 3. The viscosity of coarse
emulsions ranged between 22.95 and 36.05 mPa s depending on
the type of essential oil used in the formulation. Mint oil led to
more viscous emulsions, whereas thyme oil rendered the lowest
viscosity values. The characteristics of the dispersed phase have a
significant impact on the rheological properties of emulsions
(McClements, 2005; Pal, 2011; T. F. Tadros, 1994). Similarly to our
results, Bonilla et al. (2012) found that thyme oil-loaded emulsions
stabilized with chitosan showed lower viscosity values in com-
parison with emulsions containing basil essential oil. They attrib-
uted these variations to differences in the dissociation degree and
number of ionizable compounds of the different essential oils, since
biopolymeric molecules would be adsorbed at the oil droplet sur-
face and lead to changes in the effective thickening concentration
Fig. 2. z-potential of essential oil-alginate emulsions or nanoemulsions produced by
in the aqueous phase.
high shear homogenization (3400 rpm, 2 min) (HSH) or microfluidization (150 MPa, 3 In addition, a dramatic reduction of emulsion viscosity has been
cycles) (MF). observed after microfluidization, regardless of the essential oil used

Please cite this article in press as: Salvia-Trujillo, L., et al., Physicochemical characterization and antimicrobial activity of food-grade emulsions
and nanoemulsions incorporating essential oils, Food Hydrocolloids (2014), http://dx.doi.org/10.1016/j.foodhyd.2014.07.012
6 L. Salvia-Trujillo et al. / Food Hydrocolloids xxx (2014) 1e10

Fig. 3. Viscosity (mPa s) of essential oil-alginate emulsions or nanoemulsions pro- Fig. 4. Whiteness index (WI) of essential oil-alginate emulsions or nanoemulsions
duced by high shear homogenization (3400 rpm, 2 min) (HSH) or microfluidization produced by high shear homogenization (3400 rpm, 2 min) (HSH) or microfluidization
(150 MPa, 3 cycles) (MF). (150 MPa, 3 cycles) (MF).

Nevertheless, emulsions incorporating thyme essential oil exhibi-


in the formulation. Coarse emulsion viscosity values decreased ted an opposite behavior, as the whiteness index of microfluidized
down to a 60.66% after microfluidization. In this sense, nano- nanoemulsions was higher in comparison with the coarse emul-
emulsions presented viscosities ranging between 9.77 and sion. It is known that oil droplet size significantly influences the
14.95 mPa s. Furthermore, there were no significant differences in optical properties of emulsions (McClements, 2002). Large particles
viscosity among microfluidized nanoemulsions with different scatter the light more intensely than smaller ones, which causes an
essential oils. The droplet size influences the emulsion viscosity, as increase in the lightness, opacity and whiteness index of emulsions
it has been demonstrated in template oil-in-water emulsions (McClements, 2011). In general, nanoemulsions are described as
(Chanamai & McClements, 2000; Corte s-Mun~ oz, Chevalier-Lucia, & transparent systems due to their reduced droplet size (Mason,
Dumay, 2009; Derkach, 2009; Pal, 2011). However, in the present Wilking, Meleson, Chang, & Graves, 2006; McClements, 2011;
work the thickening properties of the continuous phase play a Solans et al., 2005; T. Tadros, Izquierdo, Esquena, & Solans, 2004).
fundamental role to determine the rheological behavior of the However, in spite that all coarse emulsions of the present study
nanoemulsion system. It is known that microfluidization process- exhibited a significant decrease in their average droplet size and
ing may induce conformational changes on the molecular structure size distribution after being microfluidized, not all of them became
of biopolymers, as well as a significant reduction on their molecular completely transparent (Fig. 5). In fact, only tea tree, geranium and
weight (Lagoueyte & Paquin, 1998). marjoram nanoemulsions became completely transparent after 3
Recently, Villay et al. (2011) observed a viscosity reduction on microfluidization cycles, which is consistent with the light scat-
sodium alginate solutions when they were treated with high tering measurements showing that these nanoemulsions presented
pressure homogenization at 195 MPa and several cycles. They the lowest average droplet size and narrowest size distribution
attributed this effect to a decrease in the molar mass of sodium (Table 2, Fig. 1). Moreover, some essential oils might be water-
alginate due to a depolymerization of chains caused by covalent soluble due to their polar compounds, thus being able to solubi-
bond disruption. Similarly, Harte and Venegas (2010) reported a lize in the aqueous phase and leading also to transparent systems
75e85% viscosity reduction in alginate, k-carrageenan and xanthan (Ziani et al., 2012). In line with our results, Wooster et al. (2008)
gum solutions processed with high pressure homogenization from obtained transparent nanoemulsions containing peanut oil by
1 to 5 cycles up to 300 MPa. Hence, in basis of the results observed microfluidization (100 MPa and 5 cycles), with a droplet size lower
in the present work, it can be assumed that sodium alginate may than 60 nm.
undergo molecular changes in its molecular structure during
microfluidization process, leading to a decrease in its thickening
properties and therefore in the viscosity of nanoemulsions.

3.4. Whiteness index

The type of oil used in the formulation significantly influenced


the whiteness index of coarse emulsions and nanoemulsions
(Fig. 4). Coarse emulsions containing lemongrass, thyme, geranium
or palmarosa essential oils were those presenting higher whiteness
index, with values of 47.9, 49.6, 57.8 or 51.3, respectively. On the
contrary, coarse emulsions formulated with clove, tea tree or
marjoram essential oils exhibited the lowest whiteness index, with
values of 24.4, 29.9 or 27.8, respectively. In general, micro-
fluidization significantly affected the whiteness index of mixtures
generating nanoemulsions with lower whiteness index. Nonethe-
less, the extent of whiteness index decrease after microfluidization
depended on the oil type used in the formulation. For instance,
Fig. 5. Images of essential oil-alginate emulsions or nanoemulsions produced by high
nanoemulsions incorporating lemongrass, sage or geranium shear homogenization (3400 rpm, 2 min) (1) or microfluidization (150 MPa, 3 cycles)
essential oils showed a drastic decrease in their whiteness index (2). Lemongrass (L), clove (C), tea tree (T), thyme (Th), geranium (G), marjoram (M),
from 47.9, 36.0 and 57.8 down to 32.2, 27.7 and 49.6, respectively. palmarosa (Pr), rosewood (R), sage (S), mint (Mi).

Please cite this article in press as: Salvia-Trujillo, L., et al., Physicochemical characterization and antimicrobial activity of food-grade emulsions
and nanoemulsions incorporating essential oils, Food Hydrocolloids (2014), http://dx.doi.org/10.1016/j.foodhyd.2014.07.012
L. Salvia-Trujillo et al. / Food Hydrocolloids xxx (2014) 1e10 7

Fig. 6. Survival fraction of Escherichia coli through 30 min of reaction time with essential oil-alginate emulsions or nanoemulsions produced by high shear homogenization
(3400 rpm, 2 min) (-) or microfluidization (150 MPa, 3 cycles) (:). A control with water is represented (A). The plotted lines of log S vs. reaction time (min) correspond to the data
fitted to a Weibull equation. Data shown are a mean ± standard deviation.

Please cite this article in press as: Salvia-Trujillo, L., et al., Physicochemical characterization and antimicrobial activity of food-grade emulsions
and nanoemulsions incorporating essential oils, Food Hydrocolloids (2014), http://dx.doi.org/10.1016/j.foodhyd.2014.07.012
8 L. Salvia-Trujillo et al. / Food Hydrocolloids xxx (2014) 1e10

3.5. Antimicrobial activity the coarse emulsion formulated with lemongrass essential oil pre-
sented a b of 22.30, its respective microfluidized nanoemulsion
The inactivation kinetics of emulsions and nanoemulsions exhibited a b of 6.97 (Table 3). Similarly, the b factor of the inacti-
against E. coli is shown in Fig. 6. A Weibull equation was fitted to vation kinetics of E. coli with clove oil decreased from 19.71 with the
experimental data of the survival curves, with accuracy factors near coarse emulsion to 5.97 with the nanoemulsion. In addition, a
to 1 and R2 values higher than 86 (Table 3). concavity upwards was observed in the inactivation kinetics pre-
Emulsions and nanoemulsions containing lemongrass, clove, sented by both microfluidized nanoemulsions with a lower than 1,
thyme or palmarosa essential oil showed the strongest antimicrobial whereas coarse emulsions presented inactivation curves with a
activity. Namely, lemongrass or clove-loaded emulsions exhibited values higher than 1, which implies a slow release of phenolic
3.98 ± 0.01 or 3.52 ± 0.02 log reductions after being in contact 30 min compounds at early contact times. Therefore, microfluidized nano-
with the inoculated microorganism, respectively. On the contrary, emulsions containing lemongrass or clove essential oils showed a
coarse emulsions containing tea tree, geranium, marjoram, rose- faster release of antimicrobial compounds in comparison with
wood, sage or mint essential oils presented less than 1 log reduction coarse emulsions, presenting a faster inactivation of inoculated
of E. coli after 30 min of contact time. It is known that the antimi- E. coli population. The rest of microfluidized nanoemulsions con-
crobial activity of each essential oil is due to their phenolic com- taining other essential oils did not exhibit a significant enhancement
pounds and it largely depends on their concentration as well as on of their antimicrobial potential compared to their corresponding
their chemical structure (Dorman & Deans, 2000). In general, the coarse emulsion, since the b values of the inactivation curves were
mechanism of action of essential oils against microorganisms in- slightly reduced or even were increased in some cases. It is generally
volves the interaction of phenolic compounds with the proteins assumed that nanoencapsulated lipophilic antimicrobials would be
(porins) in the cytoplasmatic membrane that can precipitate and lead able to penetrate more easily though microbial membranes thus
to leakage of ions and other cell content causing the cell breakdown leading to an improved bactericidal action. In the case of Gram-
(Nychas, Skandamis, & Tassou, 2000). Those essential oils containing negative bacteria, such as E. coli, the presence of proteins (porins)
carvacrol and thymol terpenes, such as thyme essential oil, are re- that form channels (pores) in the outer membrane allow the influx
ported to have a strong bactericidal action (Lambert, Skandamis, of hydrophilic particles, being a barrier for lipophilic compounds
Coote, & Nychas, 2001). Nevertheless, eugenol and citral, the major and also restricting the entry of compounds by size (Schweizer,
components of clove and lemongrass or rosewood essential oil, 2012; Taber, 2008). Therefore, it would be reasonable to assume
respectively (Friedman, Henika, Levin, & Mandrell, 2004), have been that the solubilization of lipophilic antimicrobials within nano-
found to inactivate a broad spectrum of microorganisms (Hammer, emulsions would lead to an increase of binding sites able to interact
Carson, & Riley, 1999). Other aromatic compounds such as linalool with porins. In agreement with our results, recent research papers
(tea tree), pinene (mint), geraniol (geranium), or borneol (sage), report an enhancement on the antimicrobial activity of nano-
which can be found in essential oils, exhibit a lower inhibitory in- emulsions containing flavor oils such as thyme oil (Chang,
fluence against bacteria (Zachariah & Leela, 2006). Therefore, the McLandsborough, & McClements, 2012), peppermint oil (Liang
final antimicrobial activity of a determined essential oil is a combi- et al., 2012), limonene (Donsì et al., 2011) or eugenol and cinna-
nation of composition and concentration of each volatile compound, maldehyde (Gomes, Moreira, & Castell-Perez, 2011). In line with our
which in turn are governed by the plant variety, growing conditions results, Buranasuksombat, Kwon, Turner, and Bhandari (2011) found
and method of extraction of the essential oil. that emulsion droplet size had no influence on the antimicrobial
In our experiments it was observed that microfluidization of properties of lemon myrtle oil. These apparently controversial re-
coarse emulsions significantly affected their antimicrobial activity, sults indicate that a significant droplet size reduction does not
since nanoemulsions exhibited an increased and faster inactivation necessarily imply an increase in the functionality of essential oil
of E. coli. This behavior was clearly noticeable in the case of nanoemulsions, as their antimicrobial activity might be given by the
lemongrass or clove oil-loaded nanoemulsions. In particular, while whole delivery system and its particular ability to react with

Table 3
Estimated parameters of Weibull distribution function proposed to describe the inactivation kinetics of Escherichia coli by essential oil-alginate emulsions and nanoemulsions
processed by high shear homogenization (HSH) or microfluidization (MF).

Essential oil Homogenization b a R2 (%) Af


system

Lemongrass HSH 22.30 ± 0.66 4.68 ± 0.48 99.89 1.29


MF 6.97 ± 0.84 0.98 ± 0.07 97.91 1.23
Clove HSH 19.71 ± 0.38 3.00 ± 0.12 99.03 1.10
MF 5.97 ± 0.62 0.82 ± 0.07 97.63 1.16
Tea Tree HSH 35.56 ± 2.80 1.91 ± 0.70 99.40 1.35
MF 363.01 ± 135.37 0.10 ± 0.01 99.70 1.02
Thyme HSH 0.82 ± 0.22 0.23 ± 0.03 99.01 1.04
MF 1.87 ± 0.09 0.36 ± 0.01 97.99 1.06
Geranium HSH 41.56 ± 13.67 0.38 ± 0.18 96.91 1.06
MF 38.31 ± 9.75 0.28 ± 0.09 99.67 1.01
Marjoram HSH 52.45 ± 21.26 0.21 ± 0.18 97.64 1.04
MF 122.19 ± 24.48 0.10 ± 0.01 99.91 1.00
Palmarosa HSH 7.34 ± 0.03 1.02 ± 0.01 96.22 1.47
MF 4.39 ± 0.17 0.71 ± 0.08 99.66 1.03
Rosewood HSH 149.21 ± 87.16 1.02 ± 0.32 96.21 1.25
MF 72.35 ± 6.71 1.08 ± 0.11 92.93 1.29
Sage HSH 10,050.06 ± 12,587.12 0.27 ± 0.09 86.55 1.22
MF 3361.82 ± 38.27 0.34 ± 0.04 98.54 1.61
Mint HSH 38.69 ± 3.21 0.78 ± 0.01 94.94 1.17
MF 631.13 ± 30.14 0.12 ± 0.01 99.92 1.01

Values are expressed as mean ± standard deviation, b, is the scale factor; a, is the shape factor; Af, is the accuracy factor.

Please cite this article in press as: Salvia-Trujillo, L., et al., Physicochemical characterization and antimicrobial activity of food-grade emulsions
and nanoemulsions incorporating essential oils, Food Hydrocolloids (2014), http://dx.doi.org/10.1016/j.foodhyd.2014.07.012
L. Salvia-Trujillo et al. / Food Hydrocolloids xxx (2014) 1e10 9

microbial membranes. In this sense, it is also known that the Burt, S. (2004). Essential oils: their antibacterial properties and potential applica-
tions in foods e a review. International Journal of Food Microbiology, 94(3),
membrane permeability and cellular absorption is surfactant-
223e253.
dependent since the droplet charge is known to influence the effi- Chanamai, R., & McClements, D. J. (2000). Dependence of creaming and rheology of
cacy of nanoemulsified antimicrobials. Our results show that monodisperse oil-in-water emulsions on droplet size and concentration. Col-
nanoemulsions with different essential oils had significantly loids and Surfaces A: Physicochemical and Engineering Aspects, 172(1e3), 79e86.
Chang, Y., McLandsborough, L., & McClements, D. J. (2012). Physical properties and
different droplet charge, which indicates a distinct interaction be- antimicrobial efficacy of thyme oil nanoemulsions: influence of ripening in-
tween oil and surfactant or biopolymer molecules. In our case, hibitors. Journal of Agricultural and Food Chemistry, 60(48), 12056e12063.
Cortes-Mun ~ oz, M., Chevalier-Lucia, D., & Dumay, E. (2009). Characteristics of sub-
negative droplet charges could have blocked the entrance of anti-
micron emulsions prepared by ultra-high pressure homogenisation: effect of
microbial molecules though the microbial pores. Ziani et al. (2011) chilled or frozen storage. Food Hydrocolloids, 23(3), 640e654.
reported an antagonistic effect of cationic and anionic surfactants Delmas, T., Piraux, H., Couffin, A.-C., Texier, I., Vinet, F., Poulin, P., et al. (2011). How
on the antimicrobial activity of thyme oil nanoemulsions, which was to prepare and stabilize very small nanoemulsions. Langmuir, 27(5), 1683e1692.
Derkach, S. R. (2009). Rheology of emulsions. Advances in Colloid and Interface
attributed to the partition of surfactant molecules between the oil Science, 151(1e2), 1e23.
droplet and microbial surfaces. Dickinson, E. (2003). Hydrocolloids at interfaces and the influence on the properties
of dispersed systems. Food Hydrocolloids, 17(1), 25e39.
Dickinson, E. (2009). Hydrocolloids as emulsifiers and emulsion stabilizers. Food
4. Conclusions Hydrocolloids, 23(6), 1473e1482.
Djekic, L., & Primorac, M. (2008). The influence of cosurfactants and oils on the
formation of pharmaceutical microemulsions based on PEG-8 caprylic/capric
In the present work, a significant influence of microfluidization glycerides. International Journal of Pharmaceutics, 352(1), 231e239.
on the physicochemical properties of nanoemulsions with essential Donsì, F., Annunziata, M., Sessa, M., & Ferrari, G. (2011). Nanoencapsulation of
essential oils to enhance their antimicrobial activity in foods. LWT e Food Sci-
oils was evidenced. On the one hand, there was an appreciable ence and Technology, 44(9), 1908e1914.
decrease in the droplet size of emulsions after microfluidization, Dorman, H. J. D., & Deans, S. G. (2000). Antimicrobial agents from plants: anti-
down to values below 20 nm. The droplet size of nanoemulsions bacterial activity of plant volatile oils. Journal of Applied Microbiology, 88(2),
308e316.
could explain the whiteness index results, being transparent those
Ferreira, J. P., Alves, D., Neves, O., Silva, J., Gibbs, P. A., & Teixeira, P. C. (2010). Effects
nanoemulsions with the smallest droplets. Nonetheless, the optical of the components of two antimicrobial emulsions on food-borne pathogens.
properties of nanoemulsions were highly dependent on the oil type Food Control, 21(3), 227e230.
Friedman, M., Henika, P. R., Levin, C. E., & Mandrell, R. E. (2004). Antibacterial ac-
used in the formulation. On the other hand, the mechanical stress
tivities of plant essential oils and their components against Escherichia coli
during microfluidization caused a decrease in the nanoemulsions O157:H7 and Salmonella enterica in apple juice. Journal of Agricultural and Food
viscosity. This fact could be related with the stronger z-potential of Chemistry, 52(19), 6042e6048.
nanoemulsion in comparison with coarse emulsions. The antimi- Garti, N., & Leser, M. E. (2001). Emulsification properties of hydrocolloids. Polymers
for Advanced Technologies, 12(1e2), 123e135.
crobial activity of nanoemulsions was rather determined by the Gomes, C., Moreira, R. G., & Castell-Perez, E. (2011). Poly (DL-lactide-co-glycolide)
essential oil type used in the formulation than by their droplet size. (PLGA) nanoparticles with entrapped trans-cinnamaldehyde and eugenol for
Only in the case of nanoemulsions with lemongrass or clove antimicrobial delivery applications. Journal of Food Science, 76(2), N16eN24.
Håkansson, A., Tra €gårdh, C., & Bergenståhl, B. (2009). Studying the effects of
essential oil an enhanced antimicrobial activity against E. coli could adsorption, recoalescence and fragmentation in a high pressure homogenizer
be observed. These results evidence the potential advantages of using a dynamic simulation model. Food Hydrocolloids, 23(4), 1177e1183.
nanoemulsions as delivery systems of essential oils as preservatives Hammer, K. A., Carson, C. F., & Riley, T. V. (1999). Antimicrobial activity of essential
oils and other plant extracts. Journal of Applied Microbiology, 86(6), 985e990.
in foods but also indicate the need to elucidate the mechanism of Harte, F., & Venegas, R. (2010). A model for viscosity reduction in polysaccharides
interaction of lipid nanoparticles with biological membranes. subjected to high-pressure homogenization. Journal of Texture Studies, 41(1), 49e61.
Hasenhuettl, G. L., & Hartel, R. W. (2008). In G. L. Hasenhuettl, & R. W. Hartel (Eds.),
Food emulsifiers and their applications (Vol. 2). New York, NY: Springer
Acknowledgments Science þ Business Media.
Hatanaka, J., Chikamori, H., Sato, H., Uchida, S., Debari, K., Onoue, S., et al. (2010).
Physicochemical and pharmacological characterization of a-tocopherol-loaded
This study was supported by the Ministerio de Ciencia e nano-emulsion system. International Journal of Pharmaceutics, 396(1e2),188e193.
Innovacio n (Spain) throughout the project AGL2009-11475. Laura Hatanaka, J., Kimura, Y., Lai-Fu, Z., Onoue, S., & Yamada, S. (2008). Physicochemical
and pharmacokinetic characterization of water-soluble coenzyme Q10 formu-
Salvia-Trujillo thanks the Ministry of Science and Education (Spain) lations. International Journal of Pharmaceutics, 363(1e2), 112e117.
for the predoctoral fellowship. Prof. Olga Martín-Belloso thanks the He, W., Tan, Y., Tian, Z., Chen, L., Hu, F., & Wu, W. (2011). Food protein-stabilized
 Catalana de Recerca i Estudis Avançats (ICREA) for the
Institucio nanoemulsions as potential delivery systems for poorly water-soluble drugs:
preparation, in vitro characterization, and pharmacokinetics in rats. Interna-
Academia 2008 Award.
tional Journal of Nanomedicine, 6, 521e533.
Heurtault, B., Saulnier, P., Pech, B., Proust, J.-E., & Benoit, J.-P. (2003). Physico-
chemical stability of colloidal lipid particles. Biomaterials, 24(23), 4283e4300.
References Heydenreich, A. V., Westmeier, R., Pedersen, N., Poulsen, H. S., & Kristensen, H. G.
(2003). Preparation and purification of cationic solid lipid nanospheres e ef-
Adorjan, B., & Buchbauer, G. (2010). Biological properties of essential oils: an fects on particle size, physical stability and cell toxicity. International Journal of
updated review. Flavour and Fragrance Journal, 25(6), 407e426. Pharmaceutics, 254(1), 83e87.
Bakkali, F., Averbeck, S., Averbeck, D., & Idaomar, M. (2008). Biological effects of Hsu, J.-P., & Nacu, A. (2003). Behavior of soybean oil-in-water emulsion stabilized by
essential oils e a review. Food and Chemical Toxicology, 46(2), 446e475. nonionic surfactant. Journal of Colloid and Interface Science, 259(2), 374e381.
Bernardi, D. S., Pereira, T. A., Maciel, N. R., Bortoloto, J., Vieira, G. S., Oliveira, G. C., Jafari, S. M., Assadpoor, E., He, Y., & Bhandari, B. (2008). Re-coalescence of emulsion
et al. (2011). Formation and stability of oil-in-water nanoemulsions containing droplets during high-energy emulsification. Food Hydrocolloids, 22(7),
rice bran oil: in vitro and in vivo assessments. Journal of Nanobiotechnology, 44. 1191e1202.
s, L., Vargas, M., & Chiralt, A. (2012). Effect of essential oils and
Bonilla, J., Atare Jafari, S. M., He, Y., & Bhandari, B. (2007a). Optimization of nano-emulsions pro-
homogenization conditions on properties of chitosan-based films. Food Hy- duction by microfluidization. European Food Research and Technology, 225(5e6),
drocolloids, 26(1), 9e16. 733e741.
Bouwmeester, H., Dekkers, S., Noordam, M. Y., Hagens, W. I., Bulder, A. S., de Heer, C., Jafari, S. M., He, Y., & Bhandari, B. (2007b). Production of sub-micron emulsions by
et al. (2009). Review of health safety aspects of nanotechnologies in food ultrasound and microfluidization techniques. Journal of Food Engineering, 82(4),
production. Regulatory Toxicology and Pharmacology, 53(1), 52e62. 478e488.
Brar, S. K., & Verma, M. (2011). Measurement of nanoparticles by light-scattering Kralova, I., & Sjo€ blom, J. (2009). Surfactants used in food industry: a review. Journal
techniques. TrAC e Trends in Analytical Chemistry, 30(1), 4e17. of Dispersion Science and Technology, 30(9), 1363e1383.
Brud, W. S. (2010). Industrial uses of essential oils. In K. H. C. Baser, & G. Buchbauer Lagoueyte, N., & Paquin, P. (1998). Effects of microfluidization on the functional
(Eds.), Essential oils: Science, technology, and applications. Boca Raton, FL.: CRC Press. properties of xanthan gum. Food Hydrocolloids, 12(3), 365e371.
Buranasuksombat, U., Kwon, Y. J., Turner, M., & Bhandari, B. (2011). Influence of Lambert, R. J. W., Skandamis, P. N., Coote, P. J., & Nychas, G.-J. E. (2001). A study of
emulsion droplet size on antimicrobial properties. Food Science and Biotech- the minimum inhibitory concentration and mode of action of oregano essential
nology, 20(3), 793e800. oil, thymol and carvacrol. Journal of Applied Microbiology, 91(3), 453e462.

Please cite this article in press as: Salvia-Trujillo, L., et al., Physicochemical characterization and antimicrobial activity of food-grade emulsions
and nanoemulsions incorporating essential oils, Food Hydrocolloids (2014), http://dx.doi.org/10.1016/j.foodhyd.2014.07.012
10 L. Salvia-Trujillo et al. / Food Hydrocolloids xxx (2014) 1e10

Liang, R., Xu, S., Shoemaker, C. F., Li, Y., Zhong, F., & Huang, Q. (2012). Physical and Ross, T. (1996). Indices for performance evaluation of predictive models in food
antimicrobial properties of peppermint oil nanoemulsions. Journal of Agricul- microbiology. Journal of Applied Bacteriology, 81(5), 501e508.
tural and Food Chemistry, 60(30), 7548e7555. nchez-Gonza
Sa lez, L., Vargas, M., Gonza lez-Martínez, C., Chiralt, A., & Cha fer, M.
Maa, Y.-F., & Hsu, C. C. (1999). Performance of sonication and microfluidization for (2011). Use of essential oils in bioactive edible coatings: a review. Food Engi-
liquid-liquid emulsification. Pharmaceutical Development and Technology, 4(2), neering Reviews, 3(1), 1e16.
233e240. Schweizer, H. P. (2012). Understanding efflux in Gram-negative bacteria: opportu-
Mahdi Jafari, S., He, Y., & Bhandari, B. (2006). Nano-emulsion production by soni- nities for drug discovery. Expert Opinion on Drug Discovery, 7(7), 633e642.
cation and microfluidization e a comparison. International Journal of Food Seekkuarachchi, I. N., Tanaka, K., & Kumazawa, H. (2006). Formation and char-
Properties, 9(3), 475e485. aterization of submicrometer oil-in-water (O/W) emulsions, using high-
Mason, T. G., Wilking, J. N., Meleson, K., Chang, C. B., & Graves, S. M. (2006). energy emulsification. Industrial and Engineering Chemistry Research, 45(1),
Nanoemulsions: formation, structure, and physical properties. Journal of Physics 372e390.
Condensed Matter, 18(41), R635eR666. Sell, C. (2010). Chemistry of essential oils. In K. H. Can Baser, & G. Buchbauer (Eds.),
McClements, D. J. (2002). Colloidal basis of emulsion color. Current Opinion in Essential oils. Science, technology and applications (p. 121). Boca Raton, FL: CRC
Colloid and Interface Science, 7(5e6), 451e455. Press.
McClements, D. J. (2005). Food emulsions. Principles, practices and techniques (Vol. 2). Solans, C., Izquierdo, P., Nolla, J., Azemar, N., & Garcia-Celma, M. J. (2005). Nano-
Boca Raton, FL: CRC Press. emulsions. Current Opinion in Colloid and Interface Science, 10(3e4), 102e110.
McClements, D. J. (2011). Edible nanoemulsions: fabrication, properties, and func- Taber, H. (2008). Antibiotic permeability. In Bacterial resistance to antimicrobials (p.
tional performance. Soft Matter, 7(6), 2297e2316. 169). Boca Raton, FL: CRC Press.
Menard, N., Tsapis, N., Poirier, C., Arnauld, T., Moine, L., Lefoulon, F., et al. (2011). Tadros, T. F. (1994). Fundamental principles of emulsion rheology and their appli-
Physicochemical characterization and toxicity evaluation of steroid-based sur- cations. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 91(C),
factants designed for solubilization of poorly soluble drugs. European Journal of 39e55.
Pharmaceutical Sciences, 44(5), 595e601. Tadros, T., Izquierdo, P., Esquena, J., & Solans, C. (2004). Formation and stability of
Nambam, J. S., & Philip, J. (2012). Competitive adsorption of polymer and surfactant nano-emulsions. Advances in Colloid and Interface Science, 108e109, 303e318.
at a liquid droplet interface and its effect on flocculation of emulsion. Journal of Tajkarimi, M. M., Ibrahim, S. A., & Cliver, D. O. (2010). Antimicrobial herb and spice
Colloid and Interface Science, 366(1), 88e95. compounds in food. Food Control, 21(9), 1199e1218.
Nychas, G.-J. E., Skandamis, P. N., & Tassou, C. C. (2000). Antimicrobials from herbs Vargas, M., Cha fer, M., Albors, A., Chiralt, A., & Gonza lez-Martínez, C. (2008).
and spices. In Natural antimicrobials for the minimal processing of foods (p. 176). Physicochemical and sensory characteristics of yoghurt produced from mix-
Boca Raton, FL: CRC Press. tures of cows' and goats' milk. International Dairy Journal, 18(12), 1146e1152.
Pal, R. (2011). Rheology of simple and multiple emulsions. Current Opinion in Colloid Villay, A., Lakkis de Filippis, F., Picton, L., Le Cerf, D., Vial, C., & Michaud, P. (2011).
and Interface Science, 16(1), 41e60. Comparison of polysaccharide degradations by dynamic high-pressure ho-
Peleg, M. (2006). In M. Peleg (Ed.), Advanced quantitative microbiology for foods and mogenization. Food Hydrocolloids, 27(2), 278e286.
biosystems. Models for predicting growth and inactivation. Boca Raton, FL: CRC Wooster, T. J., Golding, M., & Sanguansri, P. (2008). Impact of oil type on nano-
Press. emulsion formation and ostwald ripening stability. Langmuir, 24(22),
Qian, C., & McClements, D. J. (2011). Formation of nanoemulsions stabilized by 12758e12765.
model food-grade emulsifiers using high-pressure homogenization: factors Yang, J. S., Jiang, B., He, W., & Xia, Y. M. (2012). Hydrophobically modified alginate
affecting particle size. Food Hydrocolloids, 25(5), 1000e1008. for emulsion of oil in water. Carbohydrate Polymers, 87(2), 1503e1506.
Rao, J., & McClements, D. J. (2011). Formation of flavor oil microemulsions, nano- Zachariah, T. J., & Leela, N. K. (2006). Volatiles from herbs and spices. In Handbook of
emulsions and emulsions: influence of composition and preparation method. herbs and spices (Vol. 3, p. 177). Boca Raton, FL: CRC Press.
Journal of Agricultural and Food Chemistry, 59(9), 5026e5035. Ziani, K., Chang, Y., McLandsborough, L., & McClements, D. J. (2011). Influence of
Rao, J., & McClements, D. J. (2012). Lemon oil solubilization in mixed surfactant surfactant charge on antimicrobial efficacy of surfactant-stabilized thyme oil
solutions: rationalizing microemulsion & nanoemulsion formation. Food Hy- nanoemulsions. Journal of Agricultural and Food Chemistry, 59(11), 6247e6255.
drocolloids, 26(1), 268e276. Ziani, K., Fang, Y., & McClements, D. J. (2012). Fabrication and stability of colloidal
Rashidi, L., & Khosravi-Darani, K. (2011). The applications of nanotechnology in food delivery systems for flavor oils: effect of composition and storage conditions.
industry. Critical Reviews in Food Science and Nutrition, 51(8), 723e730. Food Research International, 46(1), 209e216.

Please cite this article in press as: Salvia-Trujillo, L., et al., Physicochemical characterization and antimicrobial activity of food-grade emulsions
and nanoemulsions incorporating essential oils, Food Hydrocolloids (2014), http://dx.doi.org/10.1016/j.foodhyd.2014.07.012

You might also like