Sei sulla pagina 1di 10

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/234841369

Effect of thermal processes on roselle


anthocyanins encapsulated in different
polymer matrices
Article in Journal of Food Processing and Preservation April 2012
Impact Factor: 1.16 DOI: 10.1111/j.1745-4549.2011.00572.x

CITATIONS

READS

284

4 authors, including:
Zuhaili Idham

Ida Idayu Muhamad

Universiti Teknologi Malaysia

Faculty of Chemical & Energy Engineering,

14 PUBLICATIONS 38 CITATIONS

145 PUBLICATIONS 369 CITATIONS

SEE PROFILE

SEE PROFILE

Siti Hamidah Mohd-Setapar


Universiti Teknologi Malaysia
50 PUBLICATIONS 100 CITATIONS
SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Ida Idayu Muhamad


Retrieved on: 27 May 2016

Journal of Food Processing and Preservation ISSN 1745-4549

EFFECT OF THERMAL PROCESSES ON ROSELLE ANTHOCYANINS


ENCAPSULATED IN DIFFERENT POLYMER MATRICES
jfpp_572

176..184

ZUHAILI IDHAM3, IDA IDAYU MUHAMAD1,4, SITI HAMIDAH MOHD SETAPAR3 and MOHD ROJI SARMIDI2
1

Department of Bioprocess Engineering, Faculty of Chemical and Natural Resources Engineering, UTM Skudai, Johor Bahru 81310, Johor, Malaysia
Chemical Engineering Pilot Plant (CEPP), UTM Skudai, Johor Bahru, Johor, Malaysia
3
Centre of Lipid Engineering Applied Research (CLEAR), UTM Skudai, Johor Bahru, Johor, Malaysia
2

Corresponding author. TEL: 607-5535577;


FAX: 607-5536165; EMAIL:
idayu@cheme.utm.my
Received for Publication May 3, 2010
Accepted for Publication July 18, 2011
doi:10.1111/j.1745-4549.2011.00572.x

ABSTRACT
The stability of roselle anthocyanins was investigated under three different heat
treatments (60, 80 and 98C). The dry red powder from roselle calyces was produced
using spray-drying technique using four different encapsulating agents, i.e., maltodextrin, gum arabic, a combination of maltodextrin and gum arabic, and soluble
starch. The four types of matrices increased the half-life of the pigments during heat
treatment, especially at 98C (P < 0.05), compared with the roselle extract as the
control. Heat treatment results showed that the combination of maltodextrin and
gum arabic had the lowest degradation kinetic rates at 60 and 80C. Arrhenius
parameters results further confirmed that the degradation of anthocyanins was
strongly dependent on the operating temperatures during heat treatment. In this
study, all encapsulating agents, except the soluble starch, largely elongated the halflife of roselle anthocyanins compared with the nonencapsulated anthocyanins.

PRACTICAL APPLICATIONS
A spectrum of natural, red-hue primer colorants is derived from roselle plant
extracts, which is substantially free of alkaloids, enzymes, aroma or solvent residuals,
and is microencapsulated to ensure good stability and shelf life. This study measured
the thermal stability and suitability of the microencapsulated natural colorants. The
colorants are for functional use in the food, pharmaceuticals, cosmetics and other
industries.

INTRODUCTION
Anthocyanins are members of a class of nearly universal,
water-soluble, terrestrial plant pigment that can be classified
chemically as both flavonoid and phenolic. Anthocyanins
provide attractive colors such as orange, red and blue. They
are water soluble, which facilitates their incorporation into
aqueous food systems. These qualities make anthocyanins an
attractive natural food colorant. Besides their color attributes,
anthocyanins have been reported to be beneficial to health as
potent antioxidant and improves visual acuity. They have also
been observed to possess antineoplastic, radiation-protective,
vasotonic, vasoprotective, anti-inflammatory, and chemoand hepatoprotective activities (Duran et al. 2001). Therefore, they have been intensively used as functional food
ingredients and color additives.
176

As roselle (Hibiscus sabdariffa L.) is widely and easily grown


in Malaysia and tropical areas, it could be another potential
source of anthocyanins as a natural food colorant. The calyces
of roselle could contribute a brilliant red color (Duangmal
et al. 2008). Of special interest to the food industry is the
limited availability of red pigments mainly due to the prohibition of Red No. 2 (Amaranth) and the continued scrutiny of
Red No. 40 and Red No. 3 (Hallagan 1991).
The composition of roselle anthocyanins, mainly known as
mono- and diglucosides, provides limited stability against
hydration and pH changes (Jackman and Smith 1996).
Because of their high reactivity, anthocyanins are easily
degraded and form colorless or undesirable brown-colored
compounds. The stability of anthocyanins depends on a combination of environmental and chemical factors. The factors
that affect the stability of anthocyanins include temperature,

Journal of Food Processing and Preservation 36 (2012) 176184 2011 Wiley Periodicals, Inc.

Z. IDHAM ET AL.

pH, oxygen, enzymes, the presence of co-pigments and


metallic ions, ascorbic acid, sulfur dioxide, sugars, and their
degradation products (Jackman et al. 1987). Color degradation is obviously an important parameter in food processing
application such as pasteurization, cooking and boiling.
Hence, the study of food colorant with regard to the heat
temperature factor was quite important before being
applied in the food. A number of previous studies investigated the thermal stability of anthocyanins (Ahmed et al.
2004; Reyes and Cisneros-Zevallos 2007; Wang and Xu 2007;
Yang et al. 2008). However, the degradation kinetics of
encapsulated anthocyanins in the thermal treatment has not
been reported previously.
Encapsulation techniques have been widely used to
reduce interactions between food components and environmental factors such as temperature, light, moisture and
oxygen. The techniques available for microencapsulation,
both spray-drying and freeze-drying, are well-established
technologies in the food industry and are most commonly
used. Roselle powders obtained by spray-drying method
were studied by Clydesdale et al. (1979), Al-Kahtani and
Hassan (1990), Andrade and Flores (2004), and Chiou and
Langrish (2007). Some advantages of the spray-drying technique include the ability to quickly produce a dry powder
(e.g., as compared with freeze-drying method), lower cost
and the ability to control the particle size distribution. In
this study, maltodextrin (MD), gum arabic (GA) and soluble
starch (SS) were used as the encapsulating agents. They have
been used in drying many fruits such as watermelon, acai,
acerola (Righetto and Netto 2005; Quek et al. 2007; Tonon
et al. 2009) and also roselle. Anthocyanins are hydrophilic
colorants and specifically compatible with water-based gel
formulation (pectin and gum) or MD and starch as coating
molecules for polar solid matrices. The carbohydrates, such
as starch and MD, have properties that are desirable as an
encapsulating agent such as low viscosity at high solid contents and good solubility. GA, a natural plant polysaccharide
exudate of acacia, is a well-known effective wall material,
which is used for many years and still a good choice because
of its stable emulsion formation and good retention of volatiles. In this research, a combination of MD and GA was also
used as an encapsulating agent. The blend of polysaccharides alters their individual functional properties and leads
to the development of new and interesting characteristics
(Williams et al. 1992). This combination was followed by
the optimization method made by Lianfu et al. (2007).
Their result remarked that the ratio 60:40 of maltodexrin
and GA could have the highest encapsulation efficiencies of
procyanidin.
The purpose of the present study was to investigate the stability of the encapsulated anthocyanins in the different
matrixes compared with the nonencapsulated anthocyanins
at different heat temperatures.

THERMAL EFFECT ON ENCAPSULATED ANTHOCYANINS

MATERIAL AND METHOD


Materials
Roselle UMKL-1 varieties were obtained from Green Valley
Eco Village, Saleng, Kulai, Johor, Malaysia. The dried and
ground calyces were kept in a screw-capped glass bottle, protected from light and placed at -18C until used. Food-grade
MD DE 11-15, GA and SS (Kanto Chemical Co. Inc., Tokyo,
Japan) were used. Ethanol, citric acid, potassium chloride,
sodium acetate and acetonitrile were of analytical grade and
were purchased from chemical suppliers. Kuromanin chloride (cyanidin 3-glucoside) was obtained from Sigma Chemical (Selangor, Malaysia).

Extraction Procedure
Pigment extraction method was modified from Laleh et al.
(2006) by maceration of dry powdered calyces with ethanol
with 2% citric acid (50:50) for 24 h at 4C. Then, the extract
was filtered in vacuum using Wattman filter (grade 1,
Whatman, Florham Park, NJ). Plant material was
re-extracted with acidified ethanol until a faint-colored
extract was obtained. Filtrates were pooled and the plant
material was discarded. The ethanolacid was removed with
rotary vacuum evaporator at 40C. Distilled water was added
to make up the remaining aqueous extract to a known
volume. The concentration of roselle anthocyanin extract was
done by adsorption chromatography via Amberlite XAD-16
(Rohm & Haas, Frankfurt, Germany). Amberlite XAD-16 was
slurred in water for 2 h before used (Kammerer et al. 2005).
Glass column was filled with 25 mL of the polymeric resin
and the resin was washed with 50 mL of acidified water. Prior
to application of the samples, the extract was centrifuged (at
3,890 g for 10 min) to remove solid particles. Aliquots of
100 mL of the extracts were applied to the adsorbent resin.
The sample was subsequently rinsed with 125 mL of water
and 50 mL of acidified water (citric acid, pH 2.0) at a flow rate
of 3 mL/min. 200 mL of acidified ethanol was used for elution
of the pigments. Fractions of dark red pigment were collected
during sample application. Then, the solvent was removed in
the rotary evaporator at 40C under vacuum. The pigments
were redissolved in deionized distilled water until Brix was
maintained at 8. Brix was measured using Atago Refractometer (Tokyo, Japan) at 20C. After desorption of phenolic compounds, the resin was reconditioned by rinsing with 100 mL
of sodium hydroxide solution (2.5 mol/L) and 125 mL of
water. All experiments were performed in duplicate. Extracts
were stored at -20C until further treatment.

Total Anthocyanin Content


Total anthocyanin content was determined via the
pH-differential technique according to the Association of

Journal of Food Processing and Preservation 36 (2012) 176184 2011 Wiley Periodicals, Inc.

177

THERMAL EFFECT ON ENCAPSULATED ANTHOCYANINS

Z. IDHAM ET AL.

Official Analytical Chemists method proposed by Lee et al.


(2005) using two buffer systems: potassium chloride buffer,
pH 1.0 (0.025 M), and sodium acetate buffer, pH 4.5 (0.4 M).
An aliquot of the extract was transferred into a 10-mL volumetric flask and made up to 10 mL with the corresponding
buffer, and the absorbance was measured at 520 and 700 nm
by UV-VIS spectrophotometer Lambda 25 (Perkin Elmer,
Norwalk, CT). Absorbance at 700 nm is measured for haze
correction. The total anthocyanin content was calculated as
cyanindin-3-glucoside according to the following equation:

Total anthocyanins (mg L ) = A 1 M 103 D (1)


where DA = (A510 pH 1.0 - A700 pH 1.0) - (A510 pH 4.5 - A700
pH 4.5); e is the molar extinction coefficient
(26,900 L/mol/cm for cyanidin-3-glucoside); 1 is the path
length in cm; M is the molecular weight (448.8 g/mol for
cyanidin-3-glucoside); D is the dilution factor; and103 is the
conversion from g to mg.

Encapsulation Process
The carrier agents MD DE 11-15, GA, combination of both
maltodextrin and gum arabic (60:40) (MD + GA), and SS
were combined with the pigment extract (8Brix) and stirred
(40C) until all the materials were completely dissolved. A
carrier agent was added subsequently until 20% of the final
solid content was reached, and the pH was maintained at the
range 2.53. Then, the mixtures were vigorously homogenized at 14,000 rpm for 1 h at room temperature. 500 mL of
feed mixtures was prepared for further process. The resulting
feed mixture was fed into the pilot plant spray dryer at a flow
rate 9.5% and was atomized by a centrifugal atomizer. The air
temperatures at the inlet and outlet were 160 and 100C,
respectively. The prepared microcapsules were collected in a
cyclone.

Scanning Electron Microscopy

solution of powdered anthocyanins in McIlvaine buffer (pH


3) were heated in closed test tubes in a water thermostat oven.
Individual samples were withdrawn at preselected time intervals, cooled (for about 30 s) and immediately measured the
anthocyanin content. This experiment was evaluated by measuring the anthocyanin content every 30 min, 2 h, 4 h, 6 h and
8 h of heating time at 60 and 80C except for those stored at
98C that were sampled at 30, 60, 90, 180 and 210 min. The
thermal treatment differed in their incubation periods
because of the differences in the anthocyanin degradation
rates (Reyes and Cisneros-Zevallos 2007). For comparison
with nonencapsulated anthocyanins, concentrated roselle
extract with pH 3 buffer was also stored at the same
condition.

Kinetic Analysis and Arrhenius Parameter


The anthocyanin content was plotted against time (days) and
a linear regression analysis was used to determine the
adequacy of the anthocyanin degradation kinetic model. The
data of the different conditions tested were best fit by the firstorder kinetic model determined from the equation:

[Ct ] = [C0 ]e( k t )

where C0 is the anthocyanin content at 0 day (after spray


drying) and Ct is the anthocyanin content after t (time) of
stability treatment at a given temperature. Degradation
rate constants (k) were obtained from the slope of the
anthocyanin retention percentage natural log versus time by
UV/VIS spectrophotometer. First-order reaction half-life
values were calculated as (Reyes and Cisneros-Zevallos 2007)

t 1 2 = ln (2) k

(3)

The temperature dependence of k values, activation energy


(Ea) and frequency factor (A) in the heat stability experiment
was determined from the Arrhenius model (Reyes and
Cisneros-Zevallos 2007):

k = e ( Ea

Particle structures of the spray-dried powder samples were


evaluated by scanning electron microscope (SEM) (model
JSM-6390 LV; JEOL Ltd., Tokyo, Japan), as illustrated in Figure 3.2. The powder samples were attached to the SEM stubs
(10 mm) using a two-sided adhesive tape. Samples were
coated with platinum using an ion sputter coater under
vacuum. SEM was operated at an accelerating voltage of
15 kV with magnifications of 1,500, 3,500 and 7,000.

(2)

R) T

(4)

Ea/R is the slope and ln A is the intercept of the relationship


between the natural log k and (1/T) in kelvin. d values were
calculated as

D = ln (10) k

(5)

z values were calculated by plotting log D versus T (kelvin),


and Q10 values were calculated as

Q10 = 1010 z

(6)

Heat Stability
The effect of heat treatment on the stability of the encapsulated anthocyanins was investigated under three different
heating temperatures: 60, 80 and 98C. The samples of 0.04%
178

Statistical Analysis
All experimental analyses were done in duplicate. Significant
differences of the means in all the data were analyzed by

Journal of Food Processing and Preservation 36 (2012) 176184 2011 Wiley Periodicals, Inc.

Z. IDHAM ET AL.

analysis of variance at P < 0.05 level by using Design Expert


6.0.8 (Stat Ease, Inc., Minneapolis, MN). Linear regression
analysis was used to obtain the degradation rate constants in
the stability studies by using Data Analysis in Microsoft Excel
2007 (Mountain View, CA).

RESULTS AND DISCUSSION

THERMAL EFFECT ON ENCAPSULATED ANTHOCYANINS

These dents are formed by shrinkage of the particles during


drying and cooling. Similar dents were observed in the study
of Obon et al. (2009) and Tonon et al. (2009). Various wall
materials would create different wall densities to provide protection from the surrounding environment interacting with
the anthocyanins pigment in the capsule. Holes were
observed in the crust in a few particles in the anthocyanins
encapsulated with SS (Fig. 1D).

Particle Morphology
Dry microcapsules were obtained as red-purple powder
samples. Figure 1 shows the examination of SEM micrographs. The particle size of the spray-dried pigment ranged
from 600 nm to 50 mm approximately, which was presented
in the form of agglomerates. All particles showed spherical in
shape with various sizes, which is typical to materials produced by spray drying. The outer surfaces of the spray-dried
microcapsules are characterized by the presence of dents.

Thermal Stability
Thermal stability is important to determine how long the
anthocyanins can be kept or processed in the thermal temperature. Figure 2 showed that thermal stability in this study
followed the first-order kinetic reaction and showed an agreement with many literature studies such as Yang et al. (2008),
Reyes and Cisneros-Zevallos (2007), Wang and Xu (2007),
and Havlikova and Mikova (1985). The coefficients of

FIG. 1. MICROGRAPHS OF THE


MICROCAPSULES FROM THE SPRAY-DRIED
ROSELLE ANTHOCYANIN PIGMENTS IN
DIFFERENT WALL MATERIALS USING A
MAGNIFICATION OF 7,000
(A) Maltodextrin, (B) maltodextrin and gum
arabic (60:40), (C) gum arabic and (D) soluble
starch

Journal of Food Processing and Preservation 36 (2012) 176184 2011 Wiley Periodicals, Inc.

179

THERMAL EFFECT ON ENCAPSULATED ANTHOCYANINS

Z. IDHAM ET AL.

FIG. 2 FIRST-ORDER KINETIC PLOT FOR THE


RETENTION OF ANTHOCYANINS IN DIFFERENT WALL
MATERIALS AND CONTROL (ROSELLE EXTRACT)
AT (A) 60C, (B) 80C AND (C) 98C

180

Journal of Food Processing and Preservation 36 (2012) 176184 2011 Wiley Periodicals, Inc.

Z. IDHAM ET AL.

TABLE 1. DEGRADATION PARAMETERS OF


ANTHOCYANIN RETENTION IN ENCAPSULATED
AND CONTROL (ROSELLE EXTRACT) IN HEAT
STABILITY

THERMAL EFFECT ON ENCAPSULATED ANTHOCYANINS

Temperature (C)

Wall material

k sk (104/day)

R2

Half-life (h)

60

MD
MD + GA
GA
SS
C
MD
MD + GA
GA
SS
C
MD
MD + GA
GA
SS
C

0.0005 0.49a
0.0003 0.54b
0.0006 1.3c
0.0011 1.6d
0.0006 9.12c
0.0019 5.1a
0.0016 1.7b
0.0022 1.59c
0.002 2.1d
0.002 3.1d
0.006 6.14a
0.006 4.85a
0.006 3.76a
0.007 5.48b
0.008 9.51c

0.965
0.929
0.8199
0.929
0.722
0.976
0.958
0.979
0.961
0.933
0.962
0.977
0.985
0.977
0.945

23.1
38.5
19.3
10.5
19.3
6.08
7.22
5.25
5.78
5.78
1.93
1.93
1.93
1.65
1.44

80

98

Means within a column with different letters are significantly different (P < 0.05).
C, control; GA, gum arabic; MD, maltodextrin; SS, soluble starch.

determination were more than 0.9 for all cases. Gradinaru


et al. (2003) found that the thermal stability was related to the
kinetics of the degradation of anthocyanin. Simpson (1985)
suggested that thermal degradation of anthocyanins could
occur via the two mechanisms: (1) hydrolysis of the
3-glycoside linkage to form the more labile aglycone and (2)
hydrolytic opening of the pyridium ring to form a substituted
chalcone, which degrades to a brown, insoluble compound of
polyphenolic nature.
From the kinetics data (Table 1), the stability of anthocyanins was significantly influenced by the encapsulating agent.
Combination of MD and GA showed the lowest kinetic rates
at 60 and 80C. There were no significant changes in kinetic
rates in spray-dried powder encapsulated with MD, combination of MD and GA, and GA only at 98C. SS gave the lowest
protective effect of the anthocyanins toward heat treatment
with high kinetic rates at 60 and 80C. This indicated that SS at
that condition is not a suitable protector because it allowed
the anthocyanins in the core to easily leak out into water. This
can be associated with the holes that appear in the encapsulated anthocyanins using SS in the SEM micrograph. Hence,
the result shows that encapsulated anthocyanins in MD and
combination of MD and GA gained high stability in the
thermal condition.
Generally, the effect of temperature and time on the degradation kinetics of anthocyanins was significant. Kinetic constant increases with the increased in temperature used. The
effect of encapsulating agent on anthocyanins degradation
kinetic rates at different heat temperatures was significant.
The kinetic rate of the degradation of anthocyanin depends
significantly (P < 0.0001) on heat temperature used in this
study, as shown in Fig. 3.
The half-life values (Table 1) varied from 1.4 to 38.5 h for
reconstituted encapsulated samples (pH 3) at 60, 80 and 98C.

The lowest (P < 0.05) half-life was obtained at the highest


temperature, 98C, which means that in this temperature and
above, the sample with anthocyanins cannot be processed or
heated for more than 2 h. Figure 4 shows that sample encapsulated with MD and GA recorded the longest (P < 0.05) halflife value at 60 and 80C. Similar values have been reported by
Mourtzinos et al. (2008) for the degradation of roselle anthocyanins in the presence of b-cyclodextrin at 60 and 80C, i.e.,
38.5 and 7.22 h for half-life values.
It could be suggested that combination of MD and GA may
provide better protection for the encapsulated pigments
(roselle anthocyanins) due to high intermolecular interaction. According to Gaonkar and McPhearson (2006), starch
and hydrocolloid (gum) interactions are often considered
synergistic. This indicated that selected encapsulating agent
could improve the stability of anthocyanins in aqueous solution (pH 3) at several thermal temperatures.

FIG. 3. EFFECT OF TEMPERATURES ON THE KINETIC RATES OF


SPRAY-DRIED ENCAPSULATED ANTHOCYANINS
MD, maltodextrin; GA, gum arabic; SS, soluble starch; C, control

Journal of Food Processing and Preservation 36 (2012) 176184 2011 Wiley Periodicals, Inc.

181

THERMAL EFFECT ON ENCAPSULATED ANTHOCYANINS

Z. IDHAM ET AL.

FIG. 4. HALF-LIFE (H) FOR ROSELLE PIGMENTS ENCAPSULATED IN


DIFFERENT MATRICES AND DIFFERENT HEAT TEMPERATURES
MD, maltodextrin; GA, gum arabic; SS, soluble starch; C, control

Arrhenius Parameter
Heat temperature significantly influences the anthocyanins
degradation rates. The dependence of reaction rates on temperature was characterized by the activation energy value, z
value and Q10. Figure 5 shows the Arrhenius plot for anthocyanins in encapsulated samples and control (roselle extract).
The graph shows, after plotting the ln k versus 1/T, that
encapsulated anthocyanins and control were all fitted to the
Arrhenius equation. The activation energy (Ea) is the energy
needed by a system to initiate a particular process. The activation energy is often used to denote the minimum energy
needed for a specific chemical reaction to occur. The high
activation energy obtained in the control sample shows that

the dependence of nonencapsulated anthocyanins on the


temperature was high in order to form the reaction (degradation of anthocyanins). The temperature sensitivity of D
values at various temperatures is normally expressed as a
thermal resistance curve with log D value plotted against temperature. The temperature sensitivity indicator is defined as z,
a value that represents a temperature range, which results in a
10-fold change in D values. In a semi-log graph, this value
represents the temperature range between which the D value
curve passes through one logarithmic cycle. Using regression
techniques, z value can be obtained as the negative reciprocal
slope of the thermal resistance curve (regression of log D
values versus temperature) (Ramaswamy and Marcotte
2006).
Referring to Table 2, the activation energy and Q10 depend
on the encapsulating agent used for encapsulated anthocyanins. The result shows that the encapsulated anthocyanins
with combination of MD and GA gave the highest activation
energy value and Q10. This was followed by the control, MD
and GA. Encapsulated anthocyanins with SS as the wall material present the lowest activation energy value and Q10. This
result shows that combination of MD and GA needs high
energy before the anthocyanin molecules able to collide with
each other and form the degradation reaction. The energy
was induced by temperature. Higher temperature used forms
faster degradation. This was similarly observed in the degradation of anthocyanin in the control sample where the activation energy was high in both stability studies. MD and GA
alone also showed dependence on temperature for the degradation of anthocyanin reaction. SS shows the lowest activation energy and Q10. This indicated only slight changes of
anthocyanins degradation rates with different thermal temperatures. Low activation energy shows that anthocyanins
only needed a small quantity of energy in order for the degradation reaction to occur. Therefore, anthocyanins encapsulated with SS showed less heat stability.

CONCLUSION

FIG. 5. THE ARRHENIUS PLOTS FOR THE DEGRADATION OF


ENCAPSULATED ANTHOCYANINS AND CONTROL DURING THERMAL
STABILITY STUDY

182

The effect of anthocyanin content in the encapsulated


samples and control were significantly affected by heat treatment. The encapsulating agent used also influenced the degradation of anthocyanins. Combination of MD and GA
(60:40) showed the lowest kinetic rates at 60 and 80C, which
suggested that the selection of suitable encapsulating agent
could improve the stability of anthocyanins even at high temperature. Arrhenius parameter performed in the thermal
study proved that degradation of anthocyanin in reconstituted form was highly depended on temperature.
The same synergistic effects between MD and GA to encapsulate anthocyanins could be observed. It is probably due to
some structure interactions between the involved materials
that made the combination the best encapsulating agent.

Journal of Food Processing and Preservation 36 (2012) 176184 2011 Wiley Periodicals, Inc.

Z. IDHAM ET AL.

THERMAL EFFECT ON ENCAPSULATED ANTHOCYANINS

TABLE 2. THERMAL DEGRADATION PARAMETERS OF ENCAPSULATED ANTHOCYANINS AND CONTROL


Wall
material

Temperature
(C)

Activation energy
(kJ/mol)*

Frequency factor,
K0 (1/h)

D value
(day)

z value
(C)

Q10

MD

60
80
98
60
80
98
60
80
98
60
80
98
60
80
98

66.95
(0.978)

1.45E + 07

14

5.18

81.09
(0.985)

1.43E + 06

12.8

6.04

62.4
(0.986)

3.43E + 06

15

4.64

47.51
(0.854)

2.66E + 04

19

3.36

68.67
(0.947)

3.10E + 07

4,605.2
1,211.9
383.8
7,675.3
1,439.1
383.8
3,837.6
1,046.6
383.8
2,093.3
1,151.3
328.9
3,837.6
1,151.3
287.8

13

5.87

MD + GA

GA

SS

* Coefficient of determination (R2) shown in parentheses.


C, control; GA, gum arabic; MD, maltodextrin; SS, soluble starch.

However, as an encapsulating agent, MD alone also showed a


remarkable result with no significant difference from the
combination of MD and GA in the stability study. GA and SS
were not suggested to be used in encapsulating the anthocyanins. Maybe some combination of these agents could perform
better result in the stability of anthocyanins.

ACKNOWLEDGMENT
The authors gratefully thank the Chemical Engineering Pilot
Plant (CEPP), Universiti Teknologi Malaysia, Skudai, for the
financial support through bioentrepreneur program.
REFERENCES
AHMED, J., SHIVHARE, U.S. and RAGHAVAN, G.S.V. 2004.
Thermal degradation kinetics of anthocyanin and visual colour
of plum puree. Eur. Food Res. Technol. 218, 525528.
AL-KAHTANI, H.A. and HASSAN, B.H. 1990. Spray drying of
roselle (Hibiscus sabdariffa L.) extract. J. Food Sci. 55,
10731076.
ANDRADE, I. and FLORES, H. 2004. Optimization of
spray-drying of roselle extract (Hibiscus sabdariffa L.). Drying.
Proceedings of the 14th International-Drying Symposium (IDS
2004), Sao Paulo, Brazil, 597604.
CHIOU, D. and LANGRISH, T.A.G. 2007. Development and
characterisation of novel nutraceuticals with spray-drying
technology. J. Food Eng. 82, 8491.
CLYDESDALE, F.M., MAIN, J.H. and FRANCIS, F.J. 1979. Roselle
(Hibiscus sabdariffa L.) anthocyanins as colorants for beverages
and gelatin desserts. J. Food Prot. 42, 204207.
DUANGMAL, K., SAICHEUA, B. and SUEEPRASAN, S. 2008.
Colour evaluation of freeze-dried roselle extract as a natural

food colorant in a model system of a drink. LWT Food Sci.


Technol. 41, 14371445.
DURAN, E.A.P., GIUSTI, M.M., WROLSTAD, R.E. and GLORIA,
B.A. 2001. Anthocyanins from banana bracts (Musa X
paradisiaca) as potential food colorants. Food Chem. 73,
327332.
GAONKAR, A.G. and MCPHEARSON, A. 2006. Ingredient
Interactions: Effects on Food, 2nd Ed., Taylor & Francis, Boca
Raton, FL.
GRADINARU, G., BILIADERIS, C.G., KALLITHRAKA, S.,
KEFALAS, P. and GARCIA-VIGUERA, C. 2003. Thermal
stability of Hibiscus sabdariffa L. anthocyanins in solution and
in solid state: Effects of co-pigmentation and glass transition.
Food Chem. 83, 423436.
HALLAGAN, B. 1991. The use of certified food color additives in
the United States. Cereal Foods World 36, 945948.
HAVLIKOVA, L. and MIKOVA, K. 1985. Heat stability of
anthocyanins. Z Lebensm Unters Forsch 181, 427432.
JACKMAN, R.L. and SMITH, J.L. 1996. Anthocyanins and
betalains. In Natural Food Colorants, 2nd Ed., (G.A.F. Hendry
and J.D. Houghton, eds.) pp. 245309, Chapman and Hall,
London, U.K.
JACKMAN, R.L., YADA, R.Y., TUNG, M.A. and SPEERS, R.A.
1987. Anthocyanins as food colorants. J. Food Biochem. 11,
201247.
KAMMERER, D., KLJUSURIC, J.G., CARLE, R. and
SCHIEBER, A. 2005. Recovery of anthocyanins from grape
pomace extracts (Vitis vinifera L. cv. Cabernet mitos) using a
polymeric adsorber resin. Eur Food Res Technol 220, 431
437.
LALEH, G.H., FRYDOONFAR, H., HEIDARY, R., JAMEEI, R. and
ZARE, S. 2006. The effect of light, temperature, pH and species
on stability of anthocyanin pigments in four Berberis species.
Pak. J. Nutr. 5, 9092.

Journal of Food Processing and Preservation 36 (2012) 176184 2011 Wiley Periodicals, Inc.

183

THERMAL EFFECT ON ENCAPSULATED ANTHOCYANINS

Z. IDHAM ET AL.

LEE, J., DURST, R. and WROLSTAD, R. 2005. AOAC official


method 2005.02: Total monomeric anthocyanin pigment
content of fruit juices, beverages, natural colorants, and wines
by the ph differential method. In Official Methods of Analysis of
AOAC International, 18th Ed., (H. Horowitz, ed.), AOAC,
Washington, DC.
LIANFU, Z., MOU, D. and DU, Y. 2007. Procyanidins:
Extraction and microencapsulation. J. Sci. Food Agric. 87,
21922197.
MOURTZINOS, I., MAKRIS, D.P., YANNAKOPOULOU, K.,
KALOGEROPOULOS, N., MICHALI, I. and KARATHANOS,
V.T. 2008. Thermal stability of anthocyanin extract of Hibiscus
sabdariffa L. in the presence of b-cyclodextrin. J. Agric. Food
Chem. 56, 1030310310.
OBON, J.M., CASTELLAR, M.R., ALACID, M. and
FERNANDEZ LOPEZ, J.A. 2009. Production of a
red-purple food colorant from Opuntia stricta fruits
by spray-drying and its application in food model
systems. J. Food Eng. 90, 471479.
QUEK, S.Y., CHOK, N.K. and SWEDLUND, P. 2007. The
physicochemical properties of spray-dried watermelon powder.
Chem. Eng. Process. 46, 386392.
RAMASWAMY, H.S. and MARCOTTE, M. 2006. Food Processing:
Principles and Applications, Taylor & Francis Group, Boca
Raton, FL.
REYES, L.F. and CISNEROS-ZEVALLOS, L. 2007. Degradation
kinetics and colour of anthocyanins in aqueous extracts of

184

purple- and red-flesh potatoes (Solanum tuberosum L.). Food


Chem. 100, 885894.
RIGHETTO, A.M. and NETTO, F.M. 2005. Effect of encapsulating
materials on water sorption, glass transition and stability of juice
from immature acerola. Int. J. Food Prop. 8, 337346.
SIMPSON, K.L. 1985. Chemical changes in natural food
pigments. In Chemical Changes in Food During Processing (T.
Richardson and J.W. Finley, eds.) pp. 409441, Van Nostrand
Reinhold, New York, NY.
TONON, R.V., BRABET, C., PALLET, D., BRAT, P. and
HUBINGER, M.D. 2009. Physicochemical and morphological
characterization of acai (Euterpe oleraceae Mart.) powder
produced with different carrier agents. Int. J. Food Sci. Technol.
44, 19501958.
WANG, W.D. and XU, S.Y. 2007. Degradation kinetics of
anthocyanins in blackberry juice and concentrate. J. Food Eng.
82, 271275.
WILLIAMS, P.A., CLEGG, S.M., LANGDON, M.J., NISHINARI,
K. and PHILLIPS, G.O. 1992. Studies on the synergistic
interaction of konjac mannan and locust bean gum with kappa
carrageenan. In Gums and Stabilizers for the Food Industry
(G.O. Phillips, P.A. Williams and D.J. Wedlock, eds.) pp.
2092160, Oxford University Press, New York, NY.
YANG, Z., HAN, Y., GU, Z., FAN, G. and CHEN, Z. 2008. Thermal
degradation kinetics of aqueous anthocyanins and visual color
of purple corn (Zea mays L.) cob. Innov. Food Sci. Emerg.
Technol. 9, 341345.

Journal of Food Processing and Preservation 36 (2012) 176184 2011 Wiley Periodicals, Inc.

Potrebbero piacerti anche