Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Available online at www.sciencedirect.

com

Journal of Food Engineering 85 (2008) 613–624


www.elsevier.com/locate/jfoodeng

DSC thermo-oxidative stability of red chili oleoresin


microencapsulated in blended biopolymers matrices
C. Pérez-Alonso a,*, J. Cruz-Olivares a, J.F. Barrera-Pichardo a, M.E. Rodrı́guez-Huezo b,
J.G. Báez-González b, E.J. Vernon-Carter b
a
Facultad de Quı́mica, Universidad Autónoma del Estado de México, Paseo Tollocan esq. Paseo Colón S/N, CP 50120, Toluca, Estado de México, Mexico
b
Departamentos de BT e IPH, Universidad Autónoma Metropolitana-Iztapalapa, San Rafael Atlixco #186, CP 09340 Mexico, D.F., Mexico

Received 21 March 2007; received in revised form 27 August 2007; accepted 30 August 2007
Available online 7 September 2007

Abstract

Red chili oleoresin-in-water emulsions were spray-dried to obtain microcapsules with wall to core ratios of 2:1 and 4:1. The biopoly-
mers mesquite gum (MG), maltodextrin DE 10 (MD) and whey protein concentrate (WPC) were used as wall materials in two different
ratios (WPC17%–MG17%–MD66% w/w and WPC66%–MG17%–MD17% w/w). The activation energy of microcapsules stored at dif-
ferent water activities (aw) was determined by non-isothermal dynamic regime differential scanning calorimetry (DSC), and was consid-
ered as indicative of thermo-oxidative stability. Best protection against oxidation was achieved when maltodextrin predominated in the
biopolymers blends, the wall to core ratio was 4:1 and aw was 0.436. Under these conditions the activation energy required for oxidizing
the microcapsule was highest having a value of 98.5 kJ mol1.
Ó 2007 Elsevier Ltd. All rights reserved.

Keywords: Thermo-oxidative stability; Red chili oleoresin; Activation energy; Water activity; Biopolymers blends

1. Introduction tion against cancer and other degenerative diseases influ-


enced by free radical reactions (Giovannucci, 1999;
Color is an important parameter of food, as it may influ- Nguyen & Schwartz, 2000), while others have pro-vitamin
ence the acceptance or rejection of products by consumers, A effects (Delgado-Vargas & Paredes-López, 2003).
serving as a visual indicator of food quality. Colorants are Carotenoids are very sensitive to oxidative decomposition
commercially available in a wide range of colors and pre- and isomerization, so that precautions have to be taken
sentations, and may be added to foodstuffs so that they when handling them to avoid degradation and the forma-
exhibit uniform color (Newsome, 1990; Pszczola, 1998). tion of artifacts such as epoxides (Britton, 1996). The
A healthy food trend has existed among consumers that potential market for water-dispersible preparations of
has driven industry to develop healthful foods that contrib- natural carotenoids is huge and includes soft drinks, ice
ute to improve or preserve health (Pszczola, Katz, & Giese, cream, desserts, candies, soups and meat products. These
2000), but that also have better taste and consumer appeal preparations must have good stability against oxidation
(Sloan, 2006). Thus, natural colorants such as carotenoids, and isomerization, and adequate solubility properties
are a natural food ingredient choice to be used in the design (Rodrı́guez-Huezo, Pedroza-Islas, Prado-Barragan, Beri-
of new functional foods. There is epidemological evidence stain, & Vernon-Carter, 2004).
that suggests that certain carotenoids may provide protec- Microencapsulation by spray-drying is a widely
employed technique used in the food industry for protect-
*
ing foods or food ingredients against deteriorative factors
Corresponding author. Tel.: +52 7222173890; fax: +52 7222175109.
E-mail address: cepa@uaemex.mx (C. Pérez-Alonso).
and for converting liquids into functional powders that

0260-8774/$ - see front matter Ó 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jfoodeng.2007.08.020
614 C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624

may be added to other food systems (Shahidi & Han, osz-Jarszewska, 1995; Pedroza-Islas, Macı́as-Bravo, & Ver-
1993). The selection of wall material will influence the non-Carter, 2002; Smith, Cain, & Talbot, 2005). The
emulsion stability during its formation and upon the drying technique involves heating the lipids, isothermally or
process affecting the characteristics of the resulting micro- dynamically, and estimating the kinetic parameters, includ-
capsules (Ré, 1998). Wall materials for microcapsules by ing the activation energy.
spray-drying should preferably comprise biopolymers that Water activity plays an important role in the oxidation
possess good surface activity (which will affect emulsifica- of microencapsulated lipids (Beristain, Azuara, & Ver-
tion/encapsulation efficiency), are highly soluble and exhi- non-Carter, 2002a, 2002b; Rodrı́guez-Huezo et al., 2004).
bit low viscosity at high concentrations (because usually From a thermodynamic point of view, it seems that the
wall to core material is high in microcapsules, and because minimum entropy point defines the conditions of maxi-
spray-drying requires high solid concentration to be eco- mum stability of microcapsules stored under given temper-
nomically feasible), and form dense adsorbed layers ature, moisture and water activity contents (Beristain,
around the lipid being encapsulated (so as to minimise dif- Azuara, & Vernon-Carter, 2002b).
fusion of pro-oxidant agents). The objective of this work was to study the thermo-oxi-
Biopolymers used as wall material for lipid encapsula- dative stability of microencapsulated red chili oleoresin
tion by spray- drying include gum Arabic (Shaikh, Bhosale, microencapsulated in different biopolymers blends matrices
& Singhal, 2006), maltodextrins (Watanabe, Fang, Minem- and stored under different water activities, by determining
oto, Adachi, & Matsuno, 2002), whey protein concentrate the activation energy required for the oxidation process
(Hogan, McNamee, O’Riordan, & O’Sullivan, 2001) and using a dynamic regime differential scanning calorimetry
mesquite gum (Beristain & Vernon-Carter, 1994). How- technique.
ever, the expected functional properties may not result
optimum when using a single biopolymer, so it is recom- 2. Materials and methods
mended to use biopolymers blends to induce a synergistic
effect (Beristain & Vernon-Carter, 1995; Beristain, Garcı́a, 2.1. Materials
& Vernon-Carter, 1999; Kagami et al., 2003; Pedroza-Islas,
Vernon-Carter, Durán-Domı́nguez, & Trejo-Martı́nez, Whey protein concentrate with 80% protein in dry basis
1999; Sheu & Rosenberg, 1998). (WPC; Hilmar 8000, Hilmar Ingredients, Hilmer, CA,
TM

In the past, selection of wall materials for microencapsu- USA); mesquite gum (MG) hand collected in the form of
lation was done by trial and error procedures, where micro- tear drops from Prosopis laevigata trees in the Mexican
capsules were formed with pre-selected biopolymers, and State of San Luis Potosı́ and purified as indicated by
then were evaluated for specific features. This procedure Vernon-Carter et al. (1996), a very high molecular weight
is time-consuming, cumbersome and expensive. Pre-selec- neutral salt of an acidic branched polysaccharide made
tion of biopolymers for microencapsulation can be done up by a backbone of residues of (1–3) linked b-D-galactose,
qualitatively by studying the isothermal drying conditions and (1–6) side chains containing L-arabinose, L-rhamnose,
of their aqueous solutions (Adachi, 2001; Bangs & Reinec- b-D-glucuronate and 4-o-methyl-b-D-glucuronate, having
cius, 1990; Imagi, Yamashita, Adachi, & Matsuno, 1992). a small amount of protein (2.7 ± 0.06%) attached to the
Best biopolymers for protecting lipids against oxidation polysaccharide moiety, which is largely responsible for its
are those where water diffusion through the biopolymer excellent emulsifying and film forming capacity (Orozco-
matrix predominantly occurs in the falling rate drying per- Villafuerte, Cruz-Sosa, Ponce-Alquicira, & Vernon-Carter,
iod where water transport is controlled by the diffusion 2003; Vernon-Carter, Beristain, & Pedroza-Islas, 2000);
process (Matsuno & Adachi, 1993). and maltodextrin DE 10 (Maltadex 10; Complementos
TM

Báez-González, Pérez-Alonso, Beristain, Vernon-Car- Alimenticios S.A. de C.V., Naucalpan, State of México,
ter, and Vizcarra-Mendoza (2004) proposed a quantitative México) were used as wall materials. Oleoresin from red
method for pre-selecting biopolymers blends consisting on chili (Capsicum annuum) with a carotenoid concentration
analyzing the isothermal drying curves by regular regime of 59.49 g kg1 (ORC; Rodofila ; Bioquimex-Reka, S.A.
TM

theory in order to obtain an effective diffusion coefficient. de C.V., Querétaro, México) was used as the carotenoid
On the other hand, Pérez-Alonso, Báez-González, Beri- source. Commercial canola oil (CO; Toluca, State of Méx-
stain, Vernon-Carter, and Vizcarra-Mendoza (2003) ico, México) was used as lipid carrier for the red chili oleo-
reported a quantitative pre-selection of biopolymers based resin. All the water used in the experiments was bidistilled.
on estimating the activation energy by water to diffuse
through the polymeric membrane when drying isother- 2.2. Continuous phase solution preparation
mally at different temperatures in the polymeric aqueous
solutions. Biopolymer solutions were prepared in accordance to a
Oxidative stability of vegetable oils, with or without Simplex Centroid experimental design (Hare, 1974)
additives, has been studied by differential scanning calorim- (Table 1). The required amount of each biopolymer was
etry (DSC) techniques by several researchers (Gloria & dispersed in water in order to obtain a 40 wt% concentra-
Aguilera, 1998; Litwinienko, Kasprzycka-Guttman, & Jar- tion, using an Ultra-Turrax T50 Basic homogenizer
C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624 615

Table 1 keep the emulsion temperature below 30 °C. Microcapsules


Simple centroid experimental design for biopolymers blends solutions containing a theoretical total carotenoids concentration
Biopolymer(s) blends code WPC (%) MG (%) MD (%) (TCC) of 2.5 g carotenoids kg1 microcapsule were
WPC100 100 0 0 obtained by spray-drying the O/W emulsions.
MG100 0 100 0
MD100 0 0 100 2.6. Microcapsule formation
WPC50–MG50 50 50 0
WPC50–MD50 50 0 50
MG50–MD50 0 50 50 The O/W emulsions were spray-dried in a Büchi model
WPC33–MG33–MD33 33.33 33.33 33.33 190 spray- dryer (Büchi Laboratoriums Technik AG, Fla-
WPC66–MG17–MD17 66 17 17 wil, Switzerland), using an inlet air temperature of
WPC17–MG66–MD17 17 66 17 170 ± 5 °C, an outlet air temperature of 95 ± 5 °C, and
WPC17–MG17–MD66 17 17 66
an atomization pressure of 4.5 bar.
WPC: whey protein concentrate, MG: mesquite gum, MD: maltodextrin
DE 10.
2.7. Thermal stability of the pure oleoresin, wall materials,
and microcapsules

(Ika-Werke Works, Inc., Wilmington, USA). The biopoly- Microcapsule samples were put into desiccators contain-
mer solutions were left for 24 h before usage to ensure com- ing different saturated salt solutions so as to achieve differ-
plete hydration. ent water activities (aw) of 0.108 (LiCl), 0.215 (KC2H3O2),
0.318 (MgCl2), 0.436 (K2CO3) and 0.515 (MgNO3) at 35 °C
2.3. Isothermal drying (Labuza & Schmidl, 1985). This temperature was consid-
ered as appropriate for storage life conditions studies.
A TA instruments thermogravimetric analyzer (model The water activity was measured with an Aqualab water
TGA 2950, New Castle, DE, USA) was used for obtaining activity meter with temperature compensation (model ser-
the drying curves. Fifteen to thirty milligrams of each bio- ies 3TE, Decagon Devices, Inc., Pullman, WA, USA). It
polymer(s) solution(s) was placed in the equipment furnace was considered that the microcapsules achieved the aw
and dried at 50, 60, 70 and 80 °C during 90 min, using air required when two consecutive aw measurements were con-
with a relative humidity of 0.008 kg H2O kg1 dry air as stant (20 days). Samples between 4 and 5 mg of the pure
purge gas at a flow rate of 100 cm3 min1. Samples were oleoresin, wall materials, and of the microcapsules were
pre-heated from ambient temperature to test temperature placed in the furnace of a TA Instruments DSC model
using the heating ramps in order to achieve isothermal dry- 2010 (New Castle, DE, USA), and were subjected to heat-
ing conditions in the least time possible for minimizing ing rates (b) of 4, 6, 8 and 10 °C min1 from 30 to 230 °C or
mass losses (Báez-González et al., 2004). 400 °C, when required, using an oxygen flow rate of
25 cm3 min1. A blank was run using N2 in order to deter-
2.4. Effective diffusion coefficient and activation energy mine if the exothermic peaks of the samples were due to
oxidation. Measurements were done in duplicate. Thermal
The estimations of the effective diffusion (Deff) coeffi- stability of the microencapsulated carotenoids was per-
cients and of the activation energy (Ea) occurring during formed using a dynamic regime non-isothermal DSC tech-
the isothermal drying of a biopolymer(s) drop were done nique proposed by Litwinienko et al. (1995), and a detailed
as reported by Pérez-Alonso et al. (2003). A detailed proce- procedure is given in Appendix III.
dure for calculating these parameters is given in Appendix I
and II. 2.8. Statistical analyses

2.5. Preparation of the oil-in-water emulsions Activation energy data were analyzed by analysis of var-
iance and significant differences between treatments deter-
Four different types of oil-in-water emulsions (O/W) mined by Tukey’s test at P = 0.05 using the NCSS
were prepared. All had a volumetric dispersed phase (/) version 5 statistical software (Wireframe Graphics, Kays-
of 0.10, using wall to core material ratios of 2:1 and 4:1, ville, UT) (NCSS, 2001) software.
consisting of those biopolymers blends that the highest
(WPC17%–MG17%–MD66% w/w) and lowest (WPC66% 3. Results and discussion
–MG17%–MD17% w/w) activation energy during the
isothermal drying of the biopolymers blends drops. The first step was to select the best and worst biopoly-
The required amount of dispersed phase (OCR + CO) mer blends to be used as wall materials from the biopoly-
was added dropwise to the required amount of biopoly- mers list given in Table 1 for microencapsulating the red
mers solution using the Ultra-Turrax T50 homogenizer at chili oleoresin following the procedure given in Appendix
a speed of 5800 rpm during 5 min. The biopolymer solu- I. Fig. 1 shows the drying rates of selected biopolymers
tions were maintained in an iced water bath in order to solutions (pure whey protein concentrate, WPC100; pure
616 C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624

0.35
a
0.30 1E-9
WPC100
MG100
Drying rate (kg H2O/s kg d.s.)

0.25 MD100

Effective diffusion (m2/s)


WPC33-MG33-MD33
0.20 WPC17-MG17-MD66 (3) (2) (1)
1E-10

0.15

0.10 WPC100
1E-11 MG100
MD100
0.05
WPC33-MG33-MD33
WPC17-MG17-MD66
0.00
1E-12
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
-0.05
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 Moisture content (kg H2O/ kg d.s.)
Moisture content (kg H2O/kg d.s.)
b
Fig. 1. Biopolymer(s) blends drying rate as a function moisture content at
60 °C.
1E-9

Effective diffusion (m2/s)


mesquite gum, MG100; pure maltodextrin, MD100; and
biopolymers blends WPC33–MG33–MD33 and WPC17–
MG17–MD66) as a function of moisture content at 1E-10
60 °C. In general terms, all of the curves showed non-linear T=50 °C
behavior characterized by three main stages: (1) a high T=60 °C
T=70 °C
moisture content (1.5–1.3 kg H2O kg1 d.s.) stage that T=80 °C
represents the ‘‘settling down” period, during which the
drying matrix conditions come into equilibrium with the
1E-11
drying air; (2) an intermediate moisture drying content 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
region (1.3–0.9 kg H2O kg1 d.s.), where drying rate Moisture content (kg H2O/ kg d.s.)
begins at its highest value at moisture content of 1.3 kg
H2O kg1 d.s. and drops rapidly in a quasi-linear fashion Fig. 2. Biopolymer(s) blends: (a) effective diffusion as a function of
moisture content at 60 °C. Numbers in parenthesis indicate the charac-
until moisture contents of 0.9 kg H2O kg1 d.s. are
teristic drying stages mentioned in the text and (b) influence of the
reached; and (3) a low moisture content region (0.9– moisture content and different temperatures on the effective diffusion of
0.1 kg H2O kg1 d.s.), where the drying rate decreased the WPC33–MG33–MD33 biopolymers blend.
non-linearly at an increasingly slower rate as moisture con-
tent decreased (Brennan, Butters, Cowell, & Lilly, 1976;
Pérez-Alonso et al., 2003). The latter period has been asso- – moisture content curves can also be characterized by
ciated with material shrinkage (Báez-González et al., 2004). three main stages: (1) a high moisture content (1.5–
The drying rate vs. moisture content curves for all of the 1.0 kg H2O kg1 d.s.) stage, where the Deff initially had a
biopolymers solutions at other temperatures showed simi- low value at the beginning of the drying; (2) an intermedi-
lar trends. Higher drying rates occurred as the isothermal ate moisture drying content stage (1.0–0.5 kg H2O
drying temperature increased. The shape of the drying kg1d.s.), where Deff value was higher and showed less fluc-
curves shown in Fig. 1 is characteristic of the so called type tuations compared to the values depicted in the other two
I curves (Matsuno & Adachi, 1993), which are concave stages; and (3) a low moisture content stage (0.5–0.2 kg
with a positive slope and where the drying rate decreases H2O kg1 d.s.), where Deff decreased more slowly as mois-
rapidly as moisture content decreases. Materials showing ture content decreased further. Fig. 2b shows that when the
this type of drying curve are considered as good microen- blend WPC33–MG33–MD33 was dried isothermally at
capsulating agents as they tend to form fine, dense films higher temperatures Deff increased. This behaviour was typ-
as soon as the drying process commences. Water move- ical for all the other biopolymers blends.
ment through the drying matrix is dominated by diffusion The effect of temperature on the mean effective diffusion
mechanism. ðDeff Þ was adequately fitted (r > 0.90) using an Arrhenius
Fig. 2a shows Deff (considering shrinkage) as a function type correlation for all the biopolymers blend (Table 2).
of moisture content of the same biopolymers aqueous The mean activation energies for all the biopolymer(s)
solutions shown in Fig. 1 dried isothermally at 60 °C. By blends are shown in Table 2. Comparing the Ea for the
contrasting Figs. 1 and 2a, it can be observed that the Deff three pure biopolymers it was found that maltodextrin
C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624 617

Table 2
Activation energy of biopolymer(s) blends 1.0
experimental
Biopolymer(s) code Ea (kJ/mol) r SD with shrinkage
0.8 without shrinkage
WPC100 13.7h 0.987 0.05
MG100 23.9f 0.997 0.09

M=(X-Xe/X0-Xe)
MD100 30.6c 0.985 0.10 0.6
WPC50–MG50 19.3g 0.974 0.07
WPC50–MD50 28.1e 0.999 0.01
0.4
MG50–MD50 25.6e 0.941 0.06
WPC33–MG33–MD33 32.3b 0.994 0.08
WPC66–MG17–MD17 5.7i 0.956 0.03 0.2
WPC17–MG66–MD17 28.7d 0.999 0.01
WPC17–MG17–MD66 41.0a 0.995 0.05 0.0
WPC: whey protein concentrate, MG: mesquite gum, MD: maltodextrin
DE 10. Different superscripts (a–i) in columns indicate significant differ- 0 1000 2000 3000 4000 5000 6000
ences (p 6 0.05). Time (s)

Fig. 3. Experimental and calculated dimensionless moisture content as


function of time for the biopolymers blend WPC17–MG17–MD66 at
had the highest value (30.6 kJ mol1), followed by MG 60 °C.
(23.9 kJ mol1) and WPC (13.7 kJ mol1). From this point
of view, maltodextrin may be considered as the best choice
for use as wall material for microencapsulating the red chili
oleoresin, as its higher Ea indicates that it is more difficult
for a unit mass of water (and presumably oxygen) to diffuse a 3.5
through the drying polymeric membrane. There exists 187.30 °C
β =4 °C/min
ample evidence that maltodextrins provide good protection 3.0 180.90 °C
β =6 °C/min
against carotenoids degradation (Desobry, Netto, & β =8 °C/min 175.80 °C
Labuza, 1999; Wagner & Warthesen, 1995). However, 2.5 β =10 °C/min
Heat flow (mW)

maltodextrins do not possess emulsifying properties (King, 2.0


Nitrogen 168.46 °C
1995), being poor encapsulating agents when used in spray-
drying, exhibiting low lipid retention (Sankarikutty, 1.5
Sreekumar, Narayanan, & Mathew, 1988), while WPC
1.0
and MG microcapsules achieve high lipids retention (Beri-
stain & Vernon-Carter, 1994, 1995; Beristain, Garcı́a, & 0.5
Vernon-Carter, 2001, 1988; Sankarikutty et al., 1988; Ver- Nitrogen
0.0
non-Carter, Pedroza-Islas, & Beristain, 1998) besides pro-
viding good protection against oxidative processes. In -0.5
Table 2 it can be observed that the biopolymers blends con- 0 50 100 150 200 250
taining a higher proportion of maltodextrin (WPC17– Temperature (°C)
MG17–MD66 and WPC33–MG33–MD33) had slightly
higher activation energies than pure maltodextrin, so that b 2.0

the above mentioned biopolymers blends should provide β =4 °C/min


345.15 °C
1.6
a better protection against oxidation than maltodextrin β =6 °C/min
335.84 °C
on its own. We selected as wall materials for microencapsu- β =8 °C/min
lating the red chili oleoresin the biopolymers blends 1.2 β =10 °C/min 328.37 °C
Heat flow (mW)

WPC17–MG17–MD66 which had the highest Ea and


WPC66–MG17–MD17 which had the lowest Ea. 0.8 310.96 °C
It is known that some biopolymer matrices shrink upon
drying and may affect the effective diffusion (Deff) determi- 0.4

nation (Pérez-Alonso et al., 2003). Thus, the experimental


dimensionless moisture content as a function of drying 0.0

time of all the biopolymers blends in Table 1 was compared


with that obtained from the solution of Fick’s second law. -0.4

The behaviour of all of the biopolymers blends can be typ-


0 50 100 150 200 250 300 350 400 450
ified by that shown by blend WPC17–MG17–MD66 at
Temperature (°C)
60 °C (Fig. 3). It is clear from Fig. 3 that the calculated
data which consider drop shrinkage overlap the experimen- Fig. 4. Thermograms for: (a) N2 blank and pure oleoresin, and for, and
tal data, whereas the calculated data which does not con- (b) biopolymers blend WPC17–MG17–MD66 at aw = 0.108.
618 C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624

sider shrinkage show considerable deviation from the Table 3


experimental data. It is obvious that shrinkage must be Kinetic parameters for the red chili oleoresin encapsulated in the WPC17–
MG17–MD66 biopolymers blend matrix with a wall to core material ratio
taken into account for obtaining reliable Deff values. of 2:1 stored at different water activities
Fig. 4a shows the thermograms of the heat flow as a
aw ß TMP Eao Z k
function of temperature of the N2 blank and of the pure (°C min1) (K) (kJ mol1) (min1) (min1)
oleoresin. The bottom flat line corresponds to that of the
0.108 4 445 49.4e 0.0488E+06 0.18
N2 blank, showing only small fluctuations and no signifi- 6 453
cant exo- and endothermic effects. The oleoresin showed 8 461
well defined exothermal peaks that corresponded to tem- 10 471
peratures of 168.46, 175.80, 180.90 and 187.30 °C for b’s 0.215 4 461 53.9d 0.305E+06 0.21
of 4, 6, 8 and 10 °C min1, respectively, indicative that 6 468
these peaks are due to the oxidation process. Fig. 4b dis- 8 480
plays the thermogram for the WPC17–MG17–MD66 wall 10 492
material at aw of 0.108 and different values of b. The exo- 0.318 4 445 58.8c 1.22E+06 0.23
thermal peaks occurred at a much higher temperatures 6 454
(310.96, 328.37, 335.84 and 345.15 °C) than those for the 8 461
10 469
pure oleoresin for the same values of b. These results indi-
cate that the exothermic peaks shown by this wall material 0.436 4 445 68.2a 16.4E+06 0.27
6 457
8 461
10 467
a 0.515 4 449 67.2b 10.9E+06 0.26
219.17 °C
8 β=4 °C/min 6 458
β=6 °C/min 8 466
6
β=8 °C/min 10 470
β=10 °C/min
206.44 °C WPC: whey protein concentrate, MG: mesquite gum, MD: maltodextrin
Heat flow (mW)

4 DE 10. Different superscripts (a–e) in columns indicate significant differ-


199.33 °C ences (p 6 0.05).
184.52 °C
2

0
Table 4
Kinetic parameters for the red chili oleoresin encapsulated in the WPC66–
-2 MG17–MD17 biopolymers blend matrix with a wall to core material ratio
of 2:1 at different water activities
-4
aw ß TMP Eao Z k
50 100 150 200 250 (°C min1) (K) (kJ mol1) (min1) (min1)
Temperature (°C) 0.108 4 457 42.9e 0.0083E+06 0.16
6 472
10 8 479
b 10 492
β= 4 °C/min 198.14 °C
8 188.37 °C 0.215 4 457 51.6d 0.104E+06 0.19
β= 6 °C/min 6 465
β= 8 °C/min
8 479
6 β= 10 °C/min
180.07 °C 10 482
Heat flow (mW)

0.318 4 458 53.7c 0.179E+06 0.20


4 6 466
172.37 °C
8 479
2 10 484
0.436 4 458 64.1a 3.09E+06 0.24
0 6 467
8 476
10 481
-2
0.515 4 444 58.3b 1.05E+06 0.23
6 455
50 100 150 200
8 463
Temperature (°C) 10 468
Fig. 5. Microcapsules thermograms made with: (a) biopolymers blend WPC: whey protein concentrate, MG: mesquite gum, MD: maltodextrin
WPC17–MG17–MD66; and (b) biopolymers blend WPC66–MG17– DE 10. Different superscripts (a–e) in columns indicate significant differ-
MD17, both at aw = 0.108. ences (p 6 0.05).
C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624 619

are likely to be due to its degradation. The other wall mate- in Table 3. As can be seen, the activation energy increased
rial (WPC66–MG17–MD17) showed similar results as as water activity increased until a maximum was reached at
those exhibited by the WPC17–MG17–MD66 wall aw = 0.436, then decreased slightly at aw = 0.515. These
material. results indicate that a close relationship existed between
Fig. 5a and b shows thermograms of the microcapsules Ea and aw. A similar behaviour was shown by the rest of
obtained with the biopolymers blends WPC17–MG17– the microcapsules (Tables 4–6). Water content and water
MD66 and WPC66–MG17–MD17 with wall to core mate- activity can influence the activation energy. The effects
rial ratios of 2:1 at aw of 0.108, respectively. Thermograms are related to the fact that the water present at the mono-
for other wall materials and different wall to core material
ratios showed similar shaped thermograms (not shown
Table 6
here). The maximum exothermal temperatures took place
Kinetic parameters for the red chili oleoresin encapsulated in the WPC66–
between approximately 170 and 222 °C, temperatures MG17–MD17 biopolymers blend matrix with a wall to core material ratio
which were close to those used (185.14 °C) for encapsulat- of 4:1 at different water activities
ing polyunsaturated fatty acids with hydrophilic wall mate- aw ß TMP Eao Z k
rials such as ovoalbumin, caseinates and gelatin (Lin, Lin, (°C min1) (K) (kJ mol1) (min1) (min1)
& Sun, 1995; Taguchi, Iwami, Ibuki, & Kawabata, 1992). 0.108 4 456 68.1e 0.00999E+09 0.26
Ulkowski, Musialik, and Litwinienko (2005) found that 6 469
the oxidation of linoleic acid showed maximum exothermal 8 472
temperatures in the temperature range between 131 and 10 478
153 °C, and soybean lecithin between 177 and 193 °C, 0.215 4 475 74.1d 0.0237E+09 0.26
depending on the heating rate. Pedroza-Islas et al. (2002) 6 483
found maximum exothermal temperatures for cod oil 8 494
10 494
microencapsulated with biopolymers blend made up by
gum arabic, mesquite gum and maltodextrin in the temper- 0.318 4 477 83.9c 0.304E+09 0.29
ature range between 245 and 350 °C. 6 479
8 494
The kinetic parameters for the red chili oleoresin encap- 10 494
sulated in WPC17–MG17–MD66 biopolymers blend
0.436 4 482 96.9a 7.40E+09 0.34
matrix, with wall to core material ratio of 2:1 are shown
6 484
8 491
10 493
Table 5
Kinetic parameters for the red chili oleoresin encapsulated in the WPC17– 0.515 4 458 93.4b 9.71E+09 0.36
MG17–MD66 biopolymers blend matrix with a wall to core material ratio 6 465
of 4:1 at different water activities 8 472
10 473
aw ß TMP Eao Z k
(°C min1) (K) (kJ mol1) (min1) (min1) WPC: whey protein concentrate, MG: mesquite gum, MD: maltodextrin
DE 10. Different superscripts (a–e) in columns indicate significant differ-
0.108 4 477 72.0e 0.012E+09 0.25 ences (p 6 0.05).
6 484
8 494
10 499
0.215 4 477 87.2d 0.694E+09 0.31 100
Activation energy required for oxidation (kJ/mol)

6 483
8 493
90
10 494
0.318 4 480 94.2c 3.64E+09 0.33 80
6 489
8 493
70
10 494
0.436 4 461 98.5a 33.23E+09 0.37 60
6 468
8 473
WPC17-MG17-MD66 (2:1)
10 477 50 WPC66-MG17-MD17 (2:1)
WPC17-MG17-MD66 (4:1)
WPC66-MG17-MD17 (4:1)
0.515 4 480 96.5b 7.33E+09 0.34 adjustment WPC17-MG17-MD66 (2:1)
40 adjustment WPC66-MG17-MD17 (2:1)
6 489 adjustment WPC17-MG17-MD66 (4:1)
adjustment WPC66-MG17-MD17 (4:1)
8 493
10 493 0.1 0.2 0.3 0.4 0.5
WPC: whey protein concentrate, MG: mesquite gum, MD: maltodextrin Water activity (aw)
DE 10. Different superscripts (a–e) in columns indicate significant differ-
ences (p 6 0.05). Fig. 6. Microcapsules activation energy as a function of water activity.
620 C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624

Table 7
Polynomial relation between microcapsules Eao and aw
Wall to core material ratio Biopolymers blend Equation R2 SD
2:1 WPC17–MG17–MD66 Eao ¼ 44:8  63:1aw þ 547:5a2w  720:0a3w 0.913 4.6
2:1 WPC17–MG17–MD66 Eao ¼ 56:0  122:0aw þ 657:5a2w  732:6a3w 0.985 2.0
4:1 WPC66–MG17–MD17 Eao ¼ 50:3 þ 236:6aw  334:4a2w þ 96:9a3w 0.998 1.0
4:1 WPC66–MG17–MD17 Eao ¼ 82:9  258:4aw þ 1311:7a2w  1493:0a3w 0.993 2.1
WPC: whey protein concentrate, MG: mesquite gum, MD: maltodextrin DE 10.

layer was tightly bound and cannot act as an aqueous a considerably higher carotenoid degradation occurred at
phase reaction medium and/or that the rate of reaction lower wall to core material ratios.
was so slow in this water as to be negligible in terms of food Nevertheless, in this work it was established that a
storage stability Labuza (1980). It may be inferred that the strong relationship existed between the activation energy
values of the water monolayer for the red chili oleoresin for the oxidation process of microencapsulated red chili
microcapsules lay in a aw range between 0.318 and 0.515, oleoresin and the water activity at which the microcapsules
as it was in this aw range that the highest activation energy are stored as shown in Fig. 6. Furthermore, a third order
was for the oxidation process. polynomial relationship could be established between the
Wall to core material ratio influenced the activation activation energy and water activity for all the microcap-
energy for the oxidation process in all the microcapsules. sules (Table 7).
As the ratio increased from 2:1 to 4:1 the activation energy
also increased. This behaviour can be observed by compar- 4. Conclusions
ing Tables 3 and 5 for the microcapsule made with
WPC17–MG17–MD66 biopolymers blend matrix. The estimation of the activation energy for isothermally
The microcapsules that showed highest activation drying a biopolymer(s) drop provides a quantitative indica-
energy values were those made with WPC17%–MG17%– tor for choosing which biopolymer(s) blend is best for the
MD66% w/w biopolymers blend, at whichever wall to core microencapsulation by spray-drying of red chili oleoresin.
material ratio used (43.4–68.2 kJ mol1 for ratio 2:1) and Dynamic regime non-isothermal differential scanning calo-
(68.1–98.5 kJ mol1 for ratio 4:1), depending on water rimetry is an efficient technique for determining the stabil-
activity, respectively. From these results, it was confirmed ity of the microencapsulated red chili oleoresin stored
that maltodextrin played a preponderant role in the oxida- under different water activities. The activation energies
tion prevention process in the biopolymers blend as for the oxidation process of the microencapsulated red chili
reported by several other authors (Desobry et al., 1999; oleoresin depends on water activity content of the micro-
Pedroza-Islas et al., 2002; Rodrı́guez-Huezo et al., 2004; capsules, on biopolymer(s) encapsulating matrix composi-
Shahidi & Han, 1993; Watanabe et al., 2002). tion and on wall to core material ratio. Highest
Finally, it is clear that the most stable microcapsules activation energies for the oxidation process were for water
were those made with the WPC17–MG17–MD66 biopoly- activity of 0.436 and wall to core material of 4:1. A rela-
mers blend with a wall to core material ratio of 4:1, and tionship in the form of a third order polynomial could be
that highest activation energy for the oxidation process established between the activation energy for the oxidation
was at a aw = 0.436, whichever the wall to core material process and the water activity for all the microcapsules
ratio. These results agree with those found for microencap- studied.
sulated cardamom oil and orange peel oil with mesquite
gum as wall material which were reported to have maxi- Acknowledgements
mum stability against oxidation when stored at aw between
0.3 and 0.4, that corresponded to an adsorbed water mono- The authors wish to thank the partial financing of this
layer (Beristain et al., 2001, 2002b). These results suggest research to the Universidad Autónoma del Estado de Méx-
that the adsorbed water monolayer may act as a barrier ico through grant 2069/2005 and to the Consejo Nacional
against oxygen diffusion. At aw’s below monolayer values, de Ciencia y Tecnologı́a (CONACyT) through grant
oxygen may diffuse more freely through unoccupied inter- U-45992-Z.
stitial spaces. However, Rodrı́guez-Huezo et al. (2004)
reported that a blend of water and oil dispersible carote- Appendix I
noids encapsulated by spray-drying of a water-in oil-in-
water multiple emulsion stabilized by a blend of gum ara- The effective diffusion coefficients (Deff) considering
bic, mesquite gum and maltodextrin displayed an almost shrinkage and moisture dependence (Raghavan, Tulasidas,
constant good stability against carotenoids degradation Sablani, & Ramaswamy, 1995) were determined applying
in a aw range between 0 and 0.515, when the wall to core an approximate analytical solution for Fick’s second law
material in the microcapsule was approximately 4:1, but to an isotropic spherical geometry and considering a con-
C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624 621

stant moisture surface concentration (Crank, 1975) given (m2 s1), Ea (kJ mol1), R is the gas constant
by (8.314  103 kJ mol1 K1), and T is the absolute tem-
" # perature (K).
1
X  Xe 6 X 1 2 2 t
M¼ ¼ exp n p Deff 2
ð1Þ
X o  X e p2 n¼1 n2 ½RðX Þ
Appendix II
where M is the moisture ratio (dimensionless), X the moisture
content at time t (kg H2O kg1 d.s.), Xo and Xe are the initial The methodology used for determining the initial radius
and equilibrium moisture content (kg H2O kg1 d.s.), of a biopolymer drop and subsequent drop volume shrink-
respectively, and R(X) is the radius (m) of the biopolymer age (Vsh) was that proponed by Báez-González et al. (2004),
drop expressed as a function of moisture content. which basically consists on
The procedure followed is summarized below
(1) Measuring the initial density (qo) of each biopolymer
1. Obtain experimentally the isothermal drying curves at or biopolymers blends solution with a Paar digital
different temperatures. densimeter (DMA 35, Anton Paar K.G., Graz, Aus-
2. Determine the experimental moisture ratio (Mexp). tria) at 25 °C.
3. Determine (2) The volume of a biopolymer drop is given by
1=3
RðX Þ ¼ V sh Ro with shrinkage ð2Þ 4
V 0 ¼ pR30 ð6Þ
or 3
RðX Þ ¼ Ro without shrinkage ð3Þ And the density of the aqueous solutions is defined as
where Ro is the initial radius of the biopolymer drop (see m0
q0 ¼ ð7Þ
Appendix II). V0
4. Substitute Eq. (2) or (3) in Eq. (1). where m0 (kg) is the initial mass of the biopolymer
5. Assume a value of Deff using the first 50 terms in Eq. (1) drop loaded to a thermoanalyzer (TGA 2950, TA
to calculate M. Instruments, New Castle, DE, USA).
6. Compare Mexp with M and if By substituting Eq. (7) in Eq. (6) the initial radius of
jM  Mexpj > 1  1010 return to step 4, and assume a the biopolymer drop (R0) is obtained.
new Deff (3) The drop volume shrinkage (Vsh) of 40% (w/w) aque-
jM  Mexpj < 1  1010 end computation, and con- ous solutions of biopolymers as a function of mois-
sider the latest assumed value as that of Deff. ture content dried isothermally under the conditions
The activation energy (Ea) was determined calculating a mentioned in Section 2.3 by triplicate. In short the
Deff for each biopolymer blend, which in turn is tempera- procedure used was as follows: (1) a drop of biopoly-
ture dependent following an Arrhenius-type relationship. mer 40% (w/w) aqueous solution was put on a glass
The procedure followed for determining Ea was as slide; (2) micrographs of the X–Y, X–Z and Y–Z
follows planes of the drop (that exhibited an ellipsoid shape)
were taken using a; (3) the area of each plane was cal-
1. Obtain for each biopolymer treatment a plot X vs. Deff culated with an Image Analysis System software; (4)
for each isothermal drying temperature. with these areas the volume of an ellipsoid was calcu-
2. Determine for each biopolymer treatment an average lated and approximated to that of a sphere; (5) steps
effective diffusivity ðDeff Þ within the experimental tem- 1–4 were repeated as the drops were dried isother-
perature range with the following equation mally at intervals of approximately 10% moisture
R X1
Deff ðX ÞdX content decrease (determined by drop mass loss);
X
Deff ¼ 0 R X 1 ð4Þ (6) Vsh is obtained by plotting the ratio of V (the drop
X0
dX volume at a given moisture content) to V0 (the initial
drop volume) against moisture content. The experi-
where X1 is the final moisture content achieved after the
mental points are then fitted with a polynomial.
drying process (kg H2O kg1 d.s.), and Deff (X) is the
The polynomial describing Vsh was
effective diffusivity at a specific moisture content
(m2 s1). V ðX Þ
V sh ¼ ¼ A þ BX þ CX 2 ð8Þ
3. Using the following Arrhenius-type relationship V0
where A–C are coefficients of polynomial model
 
Ea
Deff ¼ Do exp  ð5Þ describing biopolymer drop shrinkage.
RT

a plot of ln Deff vs. 1/T, yields a straight line with slope The initial densities of biopolymers 40% weight aqueous
Ea/R, where Deff ðm2 s1 Þ, D0 is the Arrhenius factor solutions at 25 °C were
622 C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624

Biopolymer(s) blends code q0 (kg m3) x, from these measurements, which in this case is repre-
sented by the loss of mass due to oxidation of the sample.
WPC100 1112.9
The structural quantity is assumed to change following or-
MG100 1133.9
dinary reaction kinetics
MD100 1175.9
dx
 
WPC50–MG50 1049.9 Eao
¼ Z exp gðxÞ ð12Þ
WPC50–MD50 1144.4 dt RT
MG50–MD50 1154.9
The above assumption may generally hold. The tempera-
WPC33–MG33–MD33 1102.4
ture changes linearly
WPC66–MG17–MD17 1060.4
WPC17–MG66–MD17 1060.4 T ¼ T o þ bt ð13Þ
WPC17–MG17–MD66 1143.8 where To is the initial temperature. Combining Eqs. (12)
and (13) we obtain
and the coefficients of polynomial model describing bio- Z x
dx
Z t
polymer drop shrinkage were ¼Z expðEao =RT Þdt ð14Þ
0 gðxÞ 0
Biopolymer(s) blends code A B C GðxÞ ¼ ZH ð15Þ
WPC100 0.6925 0.4493 0.1629 Rx
where G(x) equals 0 dx=gðxÞ and H is defined as the reduced
MG100 0.6565 0.0063 0.1488 time (Ozawa, 1965). When the temperature is increased at a
MD100 0.5844 0.2011 0.0506 constant rate and the reaction barely occurs at the initial
WPC50–MG50 0.6692 0.2356 0.0099 temperature, H is given by the following equation
WPC50–MD50 0.6331 0.3221 0.0515  
MG50–MD50 0.6204 0.1078 0.0971 Eao Eao
H¼ p ð16Þ
WPC33–MG33–MD33 0.6361 0.2224 0.0136 bR RT
WPC66–MG17–MD17 0.6141 0.2074 0.0335 where the p-function is given by
WPC17–MG66–MD17 0.6598 0.1151 0.0748 Z y
WPC17–MG17–MD66 0.6642 0.3322 0.0724 expðyÞ
pðyÞ ¼  dy ð17Þ
0 y2
By using Doyle’s approximation (1962) the temperature
Appendix III integral becomes
log pðyÞ ¼ 2:315  0:4567y ð18Þ
The procedure basically consists in determining from the
Ozawa (1965) has shown that if different heating rates,
thermograms: (1) the maximum temperature (TMP) of the
b, are used the temperature TMP for a given conversion x
oxidation isotherm obtained at the four heating tempera-
may be shown, by combining Eqs. (14) and (18) and deriv-
tures (b) rates used; (2) the oxidation rate constant (ASTM,
ing Eq. (10) is obtained.
1984)
  Plotting log x against 1/TMP generally gives good
Eao straight lines of slope 0.4567(E/R). The intercept is the
k ¼ Z exp  ð9Þ
RT MP pre-exponential factor, Z, and is given by Eq. (11).

where k is the kinetic constant (s1), Z (s1) is the pre- References


exponential factor, Eao is the activation energy of the oxi-
dation process (kJ mol1), R is the universal gas constant, Adachi, S. (2001). Development of lipid-based food materials with high
functionality. Food Engineering Progress, 5, 130–133.
and TMP (K) is the maximum exothermal temperature that
ASTM (1984). Standard test method for Arrhenius kinetic constants for
occurs in the heat flow-temperature diagram; (3) the activa- thermally unstable materials E 698-79. Philadelphia, PA: ASTM.
tion energy and pre-exponential factor are computed by the Báez-González, J. G., Pérez-Alonso, C., Beristain, C. I., Vernon-Carter, E.
following equations: J., & Vizcarra-Mendoza, M. G. (2004). Effective moisture diffusivity of
biopolymer drops by regular regime theory. Food Hydrocolloids, 18,
d log b 325–333.
Eao ¼ 2:19R ð10Þ
dT 1 Bangs, W. E., & Reineccius, G. A. (1990). Characterization of selected
 MP  materials for lemon oil encapsulation by spray-drying. Journal of Food
bEao exp  RTEaoMP Science, 55, 1356–1358.
Z¼ ð11Þ Beristain, C. I., Azuara, E., & Vernon-Carter, E. J. (2002a). Thermody-
RT 2MP namic analysis of the sorption process of mesquite gum. Chemical
Engineering Communications, 189, 115–123.
where are TMP taken at the half-way interval of the b used.
Beristain, C. I., Azuara, E., & Vernon-Carter, E. J. (2002b). Effect of water
The derivation of Eqs. (10) and (11) are given by Ozawa activity on the stability to oxidation of spray-dried encapsulated
(1970). In thermal analysis the kinetic parameters cannot orange peel oil using mesquite gum (Prosopis juliflora) as wall material.
be obtained unless we can derive the structural quantity, Journal of Food Science, 67, 206–211.
C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624 623

Beristain, C. I., Garcı´a, H. S., & Vernon-Carter, E. J. (1999). Mesquite Newsome, R. L. (1990). Natural and synthetic coloring agents. In A. L.
gum (Prosopis juliflora) and maltodextrin blends as wall materials for Branen, P. M. Davidson, & S. Salminen (Eds.), Food additives
spray-dried encapsulated orange peel oil. Food Science and Technology (pp. 326–346). New York: Marcel Dekker.
International, 5, 353–356. Nguyen, M. L., & Schwartz, S. J. (2000). Lycopene. In G. A. F. Hendry &
Beristain, C. I., Garcı́a, H. S., & Vernon-Carter, E. J. (2001). Spray-dried J. D. Houghton (Eds.), Natural food colorants (pp. 153–192). London:
encapsulation of cardamom (Ellearia cardamomum) essential oil with Blackie Academic and Professional.
mesquite (Prosopis juliflora) gum. Lebensmittel Wissenschaft und Orozco-Villafuerte, J., Cruz-Sosa, F., Ponce-Alquicira, E., & Vernon-
Technologie, 34, 398–401. Carter, E. J. (2003). Mesquite gum: Fractionation and characterization
Beristain, C. I., & Vernon-Carter, E. J. (1994). Utilization of mesquite of the gum exuded from Prosopis laevigata obtained from plant tissue
(Prosopis juliflora) gum as emulsion stabilizing agent for spray- culture and from wild trees. Carbohydrate Polymers, 54, 327–333.
dried encapsulated orange peel oil. Drying Technology, 12, Ozawa, T. (1965). A new method for analyzing thermogravimetric data.
1727–1733. Bulletin of the Chemical Society of Japan, 38, 1881–1886.
Beristain, C. I., & Vernon-Carter, E. J. (1995). Studies on the interaction Ozawa, T. (1970). Kinetic analysis of derivative curves in thermal analysis.
of arabic (Acacia senegal) and mesquite (Prosopis juliflora) gum as Journal of Thermal Analysis, 2, 301–324.
emulsion stabilizing agent for spray-dried encapsulated orange peel oil. Pedroza-Islas, R., Macı́as-Bravo, S., & Vernon-Carter, E. J. (2002). Oil
Drying Technology, 13, 455–461. thermo-oxidative stability and surface oil determination of biopolymer
Brennan, J. G., Butters, J. R., Cowell, N. D., & Lilly, A. E. V. (1976). microcapsules. Revista Mexicana de Ingenierı́a Quı́mica, 1, 37–42.
Food engineering operations. London: Applied Science Publishers Pedroza-Islas, R., Vernon-Carter, E. J., Durán-Domı́nguez, C., & Trejo-
Limited. Martı́nez, S. (1999). Using biopolymer blends for shrimp feedstuff
Britton, G. (1996). Carotenoids. In G. A. F. Hendry & J. D. Houghton microencapsulation-I. Microcapsule particle size, morphology and
(Eds.), Natural food colorants (pp. 197–243). London: Blackie Aca- microstructure. Food Research International, 32, 367–374.
demic and Professional. Pérez-Alonso, C., Báez-González, J. G., Beristain, C. I., Vernon-Carter, E.
Crank, J. (1975). The mathematics of diffusion. Oxford: Clarendon Press. J., & Vizcarra-Mendoza, M. G. (2003). Estimation of the activation
Delgado-Vargas, F., & Paredes-López, O. (2003). Natural colorants for energy of carbohydrate polymers blends as selection criteria for their
foods and nutraceutical uses. Boca Raton, FL: CRC Press. use as wall material for spray-dried microcapsules. Carbohydrate
Desobry, S. A., Netto, F. M., & Labuza, T. P. (1999). Influence of Polymers, 53, 197–203.
maltodextrin systems at an equivalent 25DE on encapsulated b- Pszczola, D. E. (1998). Encapsulated ingredients: Providing the right fit.
carotene loss during storage. Journal of Food Processing and Preser- Food Technology, 52(12), 70–72.
vation, 23, 39–55. Pszczola, D. E., Katz, F., & Giese, J. (2000). Research trends in healthful
Doyle, C. D. (1962). Estimating isothermal life from thermogravimetric foods. Food Technology, 54(10), 45–46, 48, 50, 52.
data. Journal of Applied Polymer Science, 6, 639–642. Raghavan, G. S. V., Tulasidas, T. N., Sablani, S. S., & Ramaswamy, H. S.
Giovannucci, E. (1999). Tomatoes, tomato-based products, lycopene, and (1995). A method of determination of concentration dependent
cancer: Review of the epidemological literature. Journal of the National effective moisture diffusivity. Drying Technology, 13, 1477–1488.
Cancer Institute, 91, 317–331. Ré, M. I. (1998). Microencapsulation by spray-drying. Drying Technology,
Gloria, H., & Aguilera, J. M. (1998). Assessment of the quality of heated 16, 1195–1236.
oils by differential scanning calorimetry. Journal of Agricultural Food Rodrı́guez-Huezo, M. E., Pedroza-Islas, R., Prado-Barragan, L. A.,
Chemistry, 46, 1363–1368. Beristain, C. I., & Vernon-Carter, E. J. (2004). Microencapsulation by
Hare, L. B. (1974). Mixture designs applied to food formulation. Food spray-drying of multiple emulsions containing carotenoids. Journal of
Technology, 28(3), 50–62. Food Science, 69, E351–E359.
Hogan, S. A., McNamee, B. F., O’Riordan, E. D., & O’Sullivan, M. Sankarikutty, B., Sreekumar, M. M., Narayanan, C. S., & Mathew, A. G.
(2001). Microencapsulating properties of whey protein concentrate. (1988). Studies on microencapsulation of cardamom oil by spray-
Journal of Food Science, 66, 675–680. drying technique. Journal of Food Science and Technology Interna-
Imagi, J., Yamashita, D., Adachi, S., & Matsuno, R. (1992). Retarded tional, 25, 352–356.
oxidation of liquid lipids entrapped in matrices of saccharides or Shahidi, F., & Han, X. (1993). Encapsulation of food ingredients. Critical
proteins. Biotechnology and Biochemistry, 56, 1236–1240. Reviews in Food Science and Nutrition, 33, 501–547.
Kagami, Y., Sugimura, S., Fujishima, N., Matsuda, K., Kometani, T., & Shaikh, J., Bhosale, R., & Singhal, R. (2006). Microencapsulation of black
Matsumura, Y. (2003). Oxidative stability, structure, and physical pepper oleoresin. Food Chemistry, 94, 105–110.
characteristics of microcapsules formed by spray-drying of fish oil with Sheu, T. Y., & Rosenberg, M. (1998). Microstructure of microcapsules
protein and dextrin wall materials. Journal of Food Science, 68, consisting of whey protein and carbohydrates. Journal of Food Science,
2248–2255. 63, 491–494.
King, A. H. (1995). Encapsulation of Food Ingredients. In S. J. Risch & Sloan, A. E. (2006). Top 10 functional food trends. Food Technology,
G. A. Reineccius (Eds.), Encapsulation and controlled release of food 60(4), 22–24, 27–28, 30–32, 35–36, 38–39.
ingredients. ACS Symposium Series 590 (pp. 26–39). Washington D.C: Smith, K. W., Cain, F. W., & Talbot, G. (2005). Kinetic analysis of non-
American Chemical Society. isothermal differential scanning calorimetry of 1,3-dipalmitoyl-2-oleo-
Labuza, T. P. (1980). The effect of water activity on reaction kinetics of ylglycerol. Journal of Agricultural and Food Chemistry, 53, 3031–3040.
food deterioration. Food Technology, 34(4), 36–41. Taguchi, K., Iwami, K., Ibuki, F., & Kawabata, M. (1992). Oxidative
Labuza, T. P., & Schmidl, M. K. (1985). Accelerated shelf life testing of stability of sardine oil embedded in spray-dried egg white powder and
foods. Food Technology, 39(9), 57–62, 64, 134. its use for n  3 unsaturated fatty acid fortification of cookies.
Lin, C., Lin, S., & Sun, H. L. (1995). Microencapsulation of squid oil with Bioscience, Biotechnology and Biochemistry, 56, 560–563.
hydrophilic macromolecules for oxidative and thermal stabilization. Ulkowski, M., Musialik, M., & Litwinienko, G. (2005). Use of differential
Journal of Food Science, 60, 36–39. scanning calorimetry to study lipid oxidation. 1. Oxidative stability of
Litwinienko, G., Kasprzycka-Guttman, T., & Jarosz-Jarszewska, M. lecithin and linolenic acid. Journal of Agricultural and Food Chemistry,
(1995). Dynamic and isothermal DSC investigation of the kinetics of 53, 9073–9077.
thermooxidative decomposition of some edible oils. Journal of Thermal Vernon-Carter, E. J., Beristain, C. I., & Pedroza-Islas, R. (2000). Mesquite
Analysis, 45, 741–750. gum (Prosopis gum). In G. Doxastakis & V. Kiosseoglou (Eds.), Novel
Matsuno, R., & Adachi, S. (1993). Lipid encapsulation technology- macromolecules in food systems (pp. 217–238). Amsterdam: Elsevier.
techniques and applications to foods. Trends in Food Science and Vernon-Carter, E. J., Gómez, S. A., Beristain, C. I., Mosqueira, G.,
Technology, 4, 256–261. Pedroza-Islas, R., & Moreno-Terrazas, R. C. (1996). Color degrada-
624 C. Pérez-Alonso et al. / Journal of Food Engineering 85 (2008) 613–624

tion and coalescence kinetics of Aztec marigold oleoresin-in-water Wagner, L. A., & Warthesen, J. J. (1995). Stability of spray-dried
emulsions stabilized by mesquite or arabic gums and their blends. encapsulated carrot carotenes. Journal of Food Science, 60, 1048–1053.
Journal of Texture Studies, 27, 625–641. Watanabe, Y., Fang, X., Minemoto, Y., Adachi, S., & Matsuno, R.
Vernon-Carter, E. J., Pedroza-Islas, R., & Beristain, C. I. (1998). Stability (2002). Suppressive effect of saturated L-ascorbate on the oxidation of
of Capsicum annuum oleoresin-in-water emulsions containing Prosopis linoleic acid encapsulated with maltodextrin or gum arabic by spray-
and Acacia gums. Journal of Texture Studies, 29, 553–567. drying. Journal of Agricultural and Food Chemistry, 50, 3984–3987.

You might also like