You are on page 1of 8

Journal of Food Engineering 178 (2016) 12e19

Contents lists available at ScienceDirect

Journal of Food Engineering


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

Effect of blanching on enzyme activity, color changes, anthocyanin


stability and extractability of mangosteen pericarp: A kinetic study
Mahsa Ziabakhsh Deylami a, *, Russly Abdul Rahman a, b, **, Chin Ping Tan a,
Jamilah Bakar a, Lasekan Olusegun a
a
Department of Food Technology, Faculty of Food Science and Technology, Universiti Putra Malaysia, 43400 Serdang, Selangor D.E., Malaysia
b
Department of Process and Food Engineering, Faculty of Engineering, Universiti Putra Malaysia, 43400 Serdang, Selangor D.E., Malaysia

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

Article history: Mangosteen pericarp is a rich source of anthocyanins. However, the high polyphenol oxidase (PPO)
Received 23 August 2015 activity challenges potential applications of pericarp as natural colorant. The kinetics of PPO inactivation,
Received in revised form anthocyanin loss and color changes were determined over a temperature range of 60e100 ! C. First-order
30 December 2015
kinetic model provided the best prediction of the PPO inactivation (R2 " 0.977), while anthocyanin loss
Accepted 1 January 2016
Available online 5 January 2016
was described by a Weibull kinetic model (R2 " 0.969). The activation energies of PPO inactivation,
anthocyanin loss and total color changes were 43.11, 57.66 and 18.86 kJ/mol, respectively. Anthocyanin
content was the most sensitive parameter towards temperature changes, suggesting the importance of
Keywords:
Mangosteen pericarp
its monitoring as a quality parameter during thermal processing. Blanching enhanced the efficiency of
Anthocyanin anthocyanin extraction. Gathering the quantitative information on the changes of PPO activity as well as
Polyphenol oxidase quality characteristics during blanching of mangosteen pericarp is important in order to design a proper
Kinetics pre-processing condition.
Weibull model © 2016 Elsevier Ltd. All rights reserved.
First order

1. Introduction syrup. While, the pericarp composes 70% of fruit and is considered
as agriculture waste (Chiste ! et al., 2010). It has high anthocyanin
Natural colorants are considered safer than synthetic colors. concentration of 182.4e423.5 mg/100 gr (Palapol et al., 2009).
They are highly demanded in the fast-growing functional and Based on availability and anthocyanin concentration, mangosteen
natural markets. Anthocyanins are interesting natural red colorants pericarp is a promising source of natural colorant. Its major
because of their attractive coloring property, high antioxidant ac- anthocyanin is cyanidin 3-sophoroside followed by cyanidin 3-
tivity and potential health effects (He and Giusti, 2010). Natural glucoside (Palapol et al., 2009) and pelargonidin 3-glucoside at
colorants are more expensive than synthetic ones. Selection of a lower concentrations (Zarena and Udaya Sankar, 2012). Neverthe-
source rich in pigment is a way to improve yield and produce a cost- less, the rapid browning of pericarp after it is cut or crushed limits
effective natural color. The major commercial sources of anthocy- its use as a red colorant. Enzymatic browning is the main reason for
anins are grape and grape skin, with anthocyanin concentration of this discoloration.
30e750 mg/100 gr (Bridle and Timberlake, 1997) and 154.6 mg/ Polyphenol oxidase (PPO) is one of the common endogenous
100 gr (Liazid et al., 2011), respectively. enzymes that contribute to anthocyanin degradation. PPO is a
Mangosteen (Garcinia Mangostana L.) is widely cultivated in copper-containing enzyme, with a di-nuclear copper center
Southeast Asia, Central America and tropical Africa (Dembitsky (Klabunde et al., 1998). This enzyme catalyzes two different re-
et al., 2011). Fruits consist of two parts, the purplish pericarp and actions in the presence of molecular oxygen: the o-hydroxylation of
the white inner pulp. The pulp is processed into juice, jam and monophenols to o-diphenols and the oxidation of o-diphenols to o-
quinones (Klabunde et al., 1998). The quinones are very reactive
electrophilic molecules that can covalently modify nucleophiles
* Corresponding author. such as anthocyanins, which results in formation of brown pig-
** Corresponding author. Department of Food Technology, Faculty of Food Science ments called melanins (Altunkaya and Go €kmen, 2012). PPO can
and Technology, Universiti Putra Malaysia, 43400 Serdang, Selangor D.E., Malaysia. react directly with anthocyanins, though they are weak substrates
E-mail addresses: mahsazd@yahoo.com (M. Ziabakhsh Deylami), russly@upm.
edu.my (R. Abdul Rahman).
for the enzyme (Kader et al., 1997), but mostly anthocyanin

http://dx.doi.org/10.1016/j.jfoodeng.2016.01.001
0260-8774/© 2016 Elsevier Ltd. All rights reserved.
M. Ziabakhsh Deylami et al. / Journal of Food Engineering 178 (2016) 12e19 13

degradation involves a co-oxidation of enzymatically generated o- using a digital portable refractometer (Atago-Master-20 M, Japan).
quinones and/or secondary oxidation products formed from TA (% citric acid) was determined by titration of 5 mL of the juice
quinone (Fang et al., 2007; Kader et al., 1997; Ruenroengklin et al., with 0.1 M NaOH. The maturity index was calculated as SSC/TA ratio
2009). According to Falguera et al. (2012), mangosteen has one of (Palapol et al., 2009). Moisture content was determined by oven
the highest PPO activity compared to other tropical fruits and the drying at 105 ! C overnight. To measure the pH of pericarp, 15 g of
enzymatic browning of fruit is strongly related to its PPO activity. pericarp was blended with 45 mL of distilled water for 2 min and
Blanching is a thermal pretreatment to inactivate oxidase en- filtered through Whatman No.1 filter paper. The pH of filtrate was
zymes via denaturation of enzyme proteins. This unit operation has measured using a pH-meter (pH-Meter Mettler-S20 SevenEasy,
dual effect on quality characteristics of product. On the one hand, USA). All assays were performed in triplicate. Fruit firmness was
the inhibition of enzyme activity stabilizes bioactive compounds determined on whole fruit using a texture analyzer (TA-XT2, Stable
during processing and storage (Brownmiller et al., 2008; Rossi et al., Micro System Ltd. Surrey, England, UK) equipped with a 2 mm
2003); furthermore, blanching causes structural changes in plant round flat probe. Color was determined at the surface of fruits using
tissue, which leads to increase in the cell wall porosity and a Minolta CR-300 portable colorimeter (illuminant D 65). The a*/b*
consequently enhances mass transfer and extraction yield of ratio was calculated as color index (Palapol et al., 2009). Three
phenolic compounds (Cipriano et al., 2015; Lin et al., 2012; measurements were done at three different locations of five fruits.
Stamatopoulos et al., 2012). On the other hand, losses of phenolic
compounds occur due to thermal degradation and/or leaching 2.3. Preparation of crude enzyme extract
during blanching (Volden et al., 2008). Through applying optimal
blanching conditions, it is possible to achieve an acceptable inac- The enzyme extracts were obtained by blending of mangosteen
tivation of enzymes while minimizing food quality degradations. pericarp with 0.1 M potassium phosphate buffer (pH 7) (1:6 v/w).
Proper design of blanching processes requires the knowledge of The buffer contained 4% (w/v) polyvinilpolypyrrolidone (PVPP)
thermal properties of product and quantitative changes of its (SigmaeAldrich, Germany) and 1% (v/v) Triton X-100 (Sigma-
enzyme activity as well as quality attributes during thermal pro- eAldrich, Germany) (Garcia-Palazon et al., 2004). PVPP is a not-
cesses (Ling et al., 2014). This information is specific for each fruit specific phenolic absorbent and Triton X-100 is a surfactant. The
depending on its species, cultivar and environmental conditions homogenate was passed through cheesecloth followed by centri-
(Gonçalves et al., 2010). fugation at 4 ! C (4323# g, 15 min), using Kubota centrifuge, model
Kinetic modeling is a useful approach for designing and opti- 5800 (Kubota Corp., Tokyo, Japan). The supernatant was collected
mizing thermal processes in order to maximize quality (Anthon as crude enzyme extract.
and Barrett, 2002; Shivhare et al., 2009). The first step is to
choose a good kinetic model. Depending on the sample, different 2.4. Enzyme activity assay
mathematical models may be employed for prediction of degra-
dation reaction rates in foods. Generally, kinetic models can be PPO activity was assayed by measuring the initial rate of in-
classified into fundamental and empirical models (Van Boekel, crease in absorbance at 410 nm (Arnnok et al., 2010) using a Thermo
2008). Arrhenius model is a semi-empirical model that Scientific GENESYS 10S UVeVis spectrophotometer. The substrate
commonly used for modeling. Since food reactions are complex in solution was composed of 1.95 mL phosphate buffer (0.1 M, pH 7.5)
nature, it is recommended to use a pure empirical model along with and 1 mL of 0.1 M catechol. The reaction was initiated by the
the Arrhenius model and choose the model with better fit (Van addition of 50 mL enzyme extract and monitored for 5 min at 5-s
Boekel, 2008). Weibull model is a pure empirical model and has a intervals. The data represent the average of triplicate assays.
great flexibility due to its non-linear nature (Corradini and Peleg, One unit of PPO activity (U) was defined as a change in absor-
2004). It was used to model enzyme inactivation (Shivhare et al., bance of 0.001 in 1 min. The specific activity was determined as U/
2009) and anthocyanin degradation (Odriozola-Serrano et al., mg protein. Protein content was determined by the Lowry method
2009). (Peterson, 1977), using bovine serum albumin (0.1e1.0 mg/mL) as
This research is the first attempt to stabilize anthocyanins of the standard. Samples were diluted so their protein content fell
mangosteen pericarp. PPO inhibition is a crucial step in this regard. within the linear range. The average protein content of samples was
Considering the necessity of blanching mangosteen pericarp and 12.05 ± 1.98 mg/mL.
the intrinsic susceptibility of anthocyanins to heat, the aim of this Residual enzyme activity in heat-treated samples is expressed
study was to analyze and quantify the effect of thermal processing as a fraction of initial activity (C0).
on inactivation of PPO, stability and extractability of anthocyanins
as well as color changes, by applying a kinetic approach, in order to Residual enzyme activity ð%Þ ¼ Ct =C0 # 100 (1)
select a proper pretreatment condition.
Where Ct and C0 were the specific enzyme activities of treated and
2. Materials and methods untreated samples, respectively.

2.1. Plant materials 2.5. Preparation of anthocyanin extracts

Fresh mangosteens (Garcinia Mangostana L.), 35 kg, of com- The extraction was performed under equal conditions for all
mercial maturity stage (a fully purple color), were obtained from a samples following the method described by (Chiste ! et al., 2010)
local market in Serdang, Selangor, Malaysia, in Jun 2012. with a slight modification. Mangosteen pericarp was made into
paste using blender. 4 g of paste was macerated with 40 mL ethanol
2.2. Physicochemical characteristics of mangosteen 95%:HCl 1.5 M (75:15, v⁄ v) in a 250 mL beaker under stirring
(300 rpm) for 12 h at 4 ! C under light-free conditions. The extract
Soluble solids content (SSC) and titratable acidity (TA) of aril was filtered through Whatman No.1 filter paper and concentrated
juice were determined as indicator of fruit maturity stage. To at 40 ! C under vacuum using a rotary evaporator (Eyela, Tokyo,
extract the juice, the white flesh with the enclosed seeds was Japan) to remove ethanol. The concentrated extract was made up to
wrapped in cheesecloth and squeezed by hand. SSC was measured 10 mL with distilled water and centrifuged (Kubota 5800; Kubota,
14 M. Ziabakhsh Deylami et al. / Journal of Food Engineering 178 (2016) 12e19

Tokyo, Japan) at 4454# g for 20 min at 4 ! C. To study thermal ! "


degradation of anthocyanins, the pH of the extract was adjusted to ln Ct=C ¼ 'k$t (4)
4.0, as it was the natural pH of mangosteen pericarp (Table 1). 0

where Co is the initial enzyme activity or anthocyanin content or


2.6. Total anthocyanin content
color value, Ct is the corresponding value at time t, k is the rate
constant (min'1), t is the heating time (min).
The total anthocyanin content was determined according to the
The half-life time (t1/2) was calculated from the following
pH- differential method (Giusti and Wrolstad, 2001), using two
equation.
buffer systems: sodium acetate 0.4 M at pH 4.5, and potassium
chloride 0.025 M at pH 1. Absorbance was read at 517 and 700 nm. t1=2 ¼ lnð2Þ=k (5)
Anthocyanin concentration was reported as milligrams cyanidin 3-
glucoside equivalent per 100 g pericarp. where k is the first-order kinetic constant.
The experimental data for anthocyanin degradation and enzyme
Total Anthocyanin Content ðmg=LÞ
inactivation were also fitted to Weibull model (Eq. (6)) (Corradini
¼ A # MW # DF # 1000=ðε # lÞ (2) and Peleg, 2004).

where A is (A517nm e A700nm) pH1.0 e (A517nm e A700nm) pH4.5. MW is Ct =C0 ¼ expðeb$t n Þ (6)
molecular weight of cyanidin 3-glucoside (449.2 g/mol), DF is
dilution factor, ε is molar absorptivity coefficient of cyanidin 3- where b is the Weibull scale parameter (min'n) and n is the shape
glucoside (26,900 L/cm/mol), l is path length (1 cm). parameter.

2.7. Isothermal heat treatment 2.10. Temperature dependency

Thermal kinetics of PPO inactivation, anthocyanins degradation The relationship between the rate constants and the tempera-
and color changes of mangosteen pericarp were studied by ture of the reaction was studied by the Arrhenius equation (Eq. (7)),
isothermal heating at five temperature levels (60, 70, 80, 90 and by performing a nonlinear regression analysis.
100 ! C) for 12 min. An aliquot of 1 mL crude enzyme/anthocyanin " !#
# $ Ea 1 1
extract was transferred to glass test tubes (100 mm length, 10.5 mm kðTÞ ¼ k Tref Exp ' (7)
i.d., 0.95 mm thick). The tubes were well sealed and put in a ther- R Tref T
mostatic shaking water bath (Memmert-SV1422, Germany). After
predetermined time intervals, the samples were cooled rapidly in where, T is the absolute temperature, Tref the reference absolute
ice water. The temperature of water bath was verified using a digital temperature, k(T) and k(Tref) are the rate constants at temperature T
thermometer (Ellab CTD-85, Emin). An untreated sample was taken and Tref, respectively, Ea is the activation energy (kJ mol'1), R is the
as control. universal gas constant (8.314 J mol K'1). The reference temperature
was the average of the range considered (i.e. Tref ¼ 80 ! C), aiming at
2.8. Water blanching more accurate parameter estimation.

For each trial, 20 g of pericarp was blanched in 6 L hot water at 2.11. Effect of blanching on anthocyanin extraction yield
preset timeetemp combinations (60e100 ! C, 12 min). Blanching
was not continued longer than 12 min as it led to an undesirable Because of the shape similarity between the curves of total
discoloration in pericarps. Before each experiment, mangosteen anthocyanin extraction yield versus blanching time and sorption
was cut in half and the pericarp was separated from the aril. The curves, we attempted to fit Peleg model (Eq. (8)).
blanched material was cooled rapidly in ice water and drained off.
Non-blanched pericarp was taken as control. t
CðtÞ ¼ C0 þ (8)
K1 þ K2 :t
2.9. Determination of thermal kinetics parameters
where, t is the blanching time (min), C(t) is anthocyanins extraction
yield after time t (mg/100 g), C0 is anthocyanin extraction yield
The rate constants (k) for zero-order (Eq. 3) and first-order (Eq.
without blanching (mg/100 g), K1 is Peleg's rate constant and K2 is
(4)) kinetic reaction were determined according to following
Peleg capacity constant.
equations:
The equation was linearized as follow (Boussetta et al., 2009),
Ct ¼ C0 ' k$t (3) t
¼ k1 þ k2 t (9)
Ct ' C0

Table 1
Physicochemical characteristics of untreated mangosteen. 2.12. Color measurement
Pericarp properties Aril properties
The visual color of anthocyanin extracts was assessed using a
Moisture content (%) 62.03 ± 3.41 SSC (! Brix) 16.4 ± 0.1
Color index (a*/b*) 3.20 ± 0.97 TA (% citric acid) 0.53 ± 0.08 Hunter Lab Colorimeter (Hunter Associates Laboratory, Inc., Vir-
Inner diameter (mm) 48.67 ± 3.67 Maturity index (SSC/TA) 31.40 ± 4.54 ginia, USA). The equipment was set to measure total transmittance
Thickness (mm) 7.94 ± 1.17 using illuminant D65, and a 10! observer. The calibration was done
pH 4.09 ± 0.02 against standard white and black tiles. The color values were
Firmness (N) 106.7 ± 38.2
expressed in terms of L* (lightness), a* (redness/greenness) and b*
M. Ziabakhsh Deylami et al. / Journal of Food Engineering 178 (2016) 12e19 15

(yellowness/blueness). Chroma (C*) is the quantitative attribute of


colorfulness. C* (Eq. (10)) and total color difference (DE) (Eq. (11))
were calculated and modeled. The average values were obtained
from five readings on each sample.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
C* ¼ a*2 þ b*2 (10)

rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
# $2 & '2 # $2
DE ¼ L*0 ' L* þ a*0 ' a* þ b*0 ' b* (11)

where, L*0, a*0 and b*0 are initial values at time zero, and L*, a* and b*
are color values at each processing time.

2.13. Statistical analysis and model evaluation

The analysis of variance (ANOVA) was used to assess the effect of


factors on the responses. Statistical significance was determined by Fig. 1. Inactivation of PPO from mangosteen pericarp during heating at 60e100 ! C.
Tukey test (5% of confidence level) using MINITAB (Version 16,
Minitab Inc., PA, USA). Non-linear regression analyses were per-
formed to calculate the kinetic parameters using SPSS statistics as physicochemical conditions like temperature and pH (Batista
software version 22 (SPSS Inc., Chicago, IL, USA). et al., 2014; Cheng et al., 2013; Terefe et al., 2010). Besides, the
The performance of models was compared by means of the initial concentration of the enzyme also should be considered
coefficient of determination (R2), chi-square (c2) (Eq. (12)) and (Agüero et al., 2008).
standard error of the mean (SEM) (Eq. (13)). The good fit is defined In order to predict thermal inactivation of mangosteen pericarp
by the higher value of R2, lowest values of c2 and SEM. PPO, Arrhenius and Weibull models were tested for their applica-
bility. As it is shown in Table 2, both models yielded high deter-
P# $2
mination coefficients (R2 " 0.966). However, first-order model had
yexp ' ypred
2
c ¼ (12) lower prediction errors so it was chosen to describe the experi-
ðm ' pÞ mental data. The monophasic inactivation of PPO shows that
despite the possible existence of two enzyme forms (labile and
P# $2 resistant), only the resistance behavior was observed. This may be
yexp ' ypred
SEM ¼ pffiffiffiffiffi (13) due to the rapid inactivation of heat-labile fraction of the enzyme at
m the very beginning of heat treatment, therefore, the observed ki-
netics are attributed to the inactivation of the heat-resistant frac-
where, yexp is the model predicted value, ypred is the experimental tion (Agüero et al., 2008; Ganjloo et al., 2011). Monophasic first-
value, p is the number of parameters and m is the number of order kinetics model has been used to model the inactivation of
experimental data points. PPO from Solanum lycocarpum (Batista et al., 2014), mushroom
(Cheng et al., 2013) and coconut water (Tan et al., 2014).
3. Results and discussion The dependence of the inactivation constants with temperature
was adequately fitted by Arrhenius equation (Table 6). The Ea value
PPO activity (Batista et al., 2014; Fang et al., 2007; Tan et al., for thermal inactivation of mangosteen pericarp PPO was 43.11 kJ/
2014) and anthocyanin content (Palapol et al., 2009) change dur- mol. This low Ea value suggests mangosteen pericarp PPO was more
ing fruit ripening. Therefor, it is important to determine the thermo-stabile than PPO from banana (155 kJ/mol) (Ünal, 2007)
maturity stage of the samples under study. Skin color is the major and mushroom (214 kJ/mol) (Cheng et al., 2013). The Ea value for
criterion to judge maturity of mangosteen. Based on the maturity thermal inactivation of PPO from S. lycocarpum (39.9 kJ/mol)
index and color index values (Table 1) the fruits used for this study (Batista et al., 2014) and Rabdosia serra leaf (52.3 kJ/mol) (Lin et al.,
were at final maturity stage (Palapol et al., 2009). 2012) were similar to that of mangosteen pericarp. Among the
oxidase enzymes from mangosteen pericarp, PPO has lower Ea and
is slightly less thermo-stable than POD with Ea of 35.06 kJ/mol
3.1. PPO inactivation
(Ziabakhsh Deylami et al., 2014).
The specific activity of untreated PPO was 18.79 ± 0.06 U/mg
protein. Normalized inactivation curves of PPO are shown in Fig. 1. 3.2. Kinetics of anthocyanins degradation
Thermal treatment significantly (p < 0.05) reduced PPO activity but
the complete inactivation was not attained. Even after 12 min of Blanching is a necessary step to stabilize anthocyanins of
heating 23e14% residual activity remained at 90 ! C and 100 ! C, mangosteen pericarp via enzyme inactivation. Due to the inherent
respectively. This high thermo-stability is not common in PPOs thermal lability of anthocyanins, it is reasonable to study their
from plants sources. The reports on high thermal resistance of PPO thermal stability under blanching conditions. The relative antho-
are not scarce though, for instance, banana PPO with remaining cyanin retention decreased significantly (p < 0.05) with increasing
activity of 80% after 30 min at 70 ! C (Yang et al., 2001), PPO from heating time and temperature, which are illustrated in Fig 2.
persimmon with residual activity of 50% after 25 min at 80 ! C Anthocyanin retention ranged from 69.05% to 89.11% after 12 min
(Navarro et al., 2014) and strawberry PPO which showed only 28% heating at temperature of 100 ! C and 60 ! C, respectively. The
inactivation after 30 min treatment at 100 ! C (Terefe et al., 2010). thermal degradation curves indicate a non-linear nature.
The variation in thermal stability of PPO depends on source, variety, The anthocyanin loss under different temperatures was fitted
environmental conditions of cultivation, treatment medium as well with first-order model and Weibull model. The corresponding
16 M. Ziabakhsh Deylami et al. / Journal of Food Engineering 178 (2016) 12e19

Table 2
The performance of the selected models to describe inactivation of PPO from mangosteen pericarp during thermal treatment.

Temperature (! C) First-order kinetic model Weibull model

K (min'1 # 10'2) t ½ (min) R2 c2 SEM b (#10'2) n (#10'1) R2 c2 (#10'2) SEM (#10'2)

60 1.93 ± 0.13 35.91 ± 2.43 0.976 0.068 0.139 1.2 ± 0.2 11.00 ± 0.85 0.987 40.84 1.477
70 5.27 ± 0.33 13.31 ± 0.85 0.972 0.648 1.059 1.8 ± 0.5 10.29 ± 1.29 0.966 7.724 0.165
80 7.59 ± 0.40 9.13 ± 0.48 0.969 1.346 2.199 2.3 ± 0.3 10.85 ± 0.64 0.992 0.125 0.220
90 11.90 ± 0.37 5.82 ± 0.18 0.995 2.804 4.579 5.3 ± 1.3 9.82 ± 1.12 0.971 0.107 0.139
100 16.12 ± 0.50 4.30 ± 0.13 0.991 5.058 8.261 14.9 ± 2.7 8.58 ± 0.88 0.972 2.766 4.517

k, rate constant; t ½, the half-life time; b, Weibull scale parameter; n the shape parameter; R2, determination coefficient; c2, chi-square; SEM, standard error of the mean.

degradation is a complex reaction. Different serial steps and various


intermediate compounds are formed during thermal degradation
of anthocyanins, depending on anthocyanin structure (Cisse et al.,
2009) as well as nature and severity of heating (Patras et al.,
2010). Therefore, one should be cautious while deriving informa-
tion on anthocyanin stability based on a single Ea value. Even
talking about anthocyanins from the same source, degradation
strongly depends on the physicochemical characteristics of the
matrix in which anthocyanins have been treated such as pH (Kırca
et al., 2007; Sui et al., 2014), solid content (Kırca et al., 2007) and
presence of other compounds such as sugars and polyphenols
(Nayak et al., 2011).

3.3. Effect of blanching on extraction yield of anthocyanins

The extraction yield of anthocyanins increased by 26.6% and


Fig. 2. Degradation of anthocyanins from mangosteen pericarp during heating at
36.5% after 12 min of treatment at 60 and 100 ! C, respectively,
60e100 ! C. compared to non-blanched pericarp. During the first 8 min, the
anthocyanin yield continuously increased, regardless of the tem-
perature. However, after 4 min the changes were not significant
kinetic parameters are presented in Table 3. The Weibull model, (p > 0.05).
with higher R2 values and lower prediction errors, provided a Despite possible thermal degradation of anthocyanins at higher
stronger fit to the data. Different mathematical models have been temperatures, the results indicated a positive effect of blanching on
employed for prediction of anthocyanin degradation reaction rates; the extractability of anthocyanins. The effect of temperature was
first-order kinetic model was used to describe thermal degradation significant (p < 0.05). The reason might be due to the effect of heat
of anthocyanins of plum puree (Ahmed et al., 2004), purple potato on softening of the pericarp tissue and inactivation of oxidase en-
(Nayak et al., 2011) and Urmu mulberry (Kara and Erçelebi, 2013); zymes, such as PPO (Cipriano et al., 2015). Heat treatment de-
while, Weibull model well described changes in anthocyanin con- natures the proteins in cell walls, which makes the membranes
tents of fresh-cut strawberries during storage (Odriozola-Serrano porous (Stamatopoulos et al., 2012). The higher porosity increases
et al., 2009). Under different temperature condition of 5e50 ! C, permeability of cell walls and improves diffusivity of the solvent,
degradation of anthocyanins extracts from mangosteen pericarp resulting in increased extraction yield (Stamatopoulos et al., 2012).
was modeled with first-order kinetics (Chiste ! et al., 2010). Nevertheless, prolonged heating caused decrease of anthocyanin
Weibull scale parameters increased with increasing tempera- extraction yield; it may be a result of degradation and/or leaching of
ture, ranging from 4.6 to 15.0 # 10'2 min'n. It shows that antho- anthocyanins.
cyanin degradation was accelerated at higher temperatures. After Interestingly, the curves of total anthocyanin concentration as a
fitting the constant rates of the Weibull model to the Arrhenius function of blanching time (Fig. 3) showed a similar shape to the
equation, the Ea of anthocyanin degradation was 57.66 kJ/mol sorption curves. This suggests that loss of anthocyanins during
Chiste! et al. (2010) reported Ea of 61.54 kJ/mol (14.7 kcal mol) for blanching due to leaching was negligible. The low level of leaching
anthocyanin extract from mangosteen pericarp at pH 1.0. Activation loss is maybe related to the textural property of mangosteen peri-
energy is a reaction characteristic. It shows the temperature carp (Table 1). The pericarp is thick and firm, therefore, cells kept
dependence of anthocyanin degradation. Anthocyanin thermal their integrity during the blanching.

Table 3
The performance of selected models to describe loss of anthocyanins in extracts of mangosteen pericarp during thermal treatment.

Temperature (! C) First-order kinetic model Weibull model

K (min'1 # 10'3) t ½ (h) R2 c2 (# 10'3) SEM(#10'3) b (#10'2) n (#10'1) R2 c2 (#10'4) SEM (#10'4)

60 4.41 ± 0.22 2.62 ± 0.25 0.950 1.08 2.21 2.7 ± 0.1 4.51 ± 0.22 0.997 0.158 0.259
70 5.90 ± 0.35 1.96 ± 0.18 0.958 1.42 2.90 4.6 ± 0.3 3.89 ± 0.29 0.983 0.281 4.59
80 9.02 ± 0.42 1.28 ± 0.21 0.939 0.576 1.18 5.2 ± 0.4 4.61 ± 0.41 0.977 0.454 0.741
90 11.01 ± 1.69 1.05 ± 0.07 0.944 2.19 4.47 9.0 ± 0.5 3.78 ± 0.28 0.982 0.108 0.177
100 11.50 ± 0.12 1.00 ± 0.12 0.966 5.22 1.07 15.0 ± 0.3 2.79 ± 0.09 0.996 0.118 0.192

For abbreviations, see Table 2.


M. Ziabakhsh Deylami et al. / Journal of Food Engineering 178 (2016) 12e19 17

Fig. 3. Effect of blanching time at 60, 80 and 100 ! C. Because of overlapping the Fig. 4. Plot of predicted values versus actual response for extraction yield of antho-
extraction plots only the results for 60, 80 and 100 ! C are presented. cyanin after blanching.

The Hunter a* value was chosen as a physical parameter to


The Peleg model has been satisfactorily applied to describe the
describe the effect of thermal treatment on red color degradation of
extraction kinetics of polyphenols from grape seeds (Buci! c-Koji!
c
anthocyanin extracts (Ahmed et al., 2004). The red color of non-
et al., 2007) and anthocyanins (Galva !n D'Alessandro et al., 2014).
blanched and thermal treated extracts was significantly (p < 0.05)
We examined a different applicability of the Peleg equation for
different. Parameter a* decreased over the entire temperature
modeling the effect of blanching conditions on the extraction yield
range used. This may be attributed to anthocyanin degradation and
of anthocyanins. All the samples were extracted under same con-
formation of browning pigments. We found a positive linear cor-
ditions and any changes in mass transfer purely came from the
relation between a* values and anthocyanin content of pericarp
thermal pretreatment.
extracts during heating (R2 ¼ 0.716e0.870). The same correlation
The calculated parameters of the Peleg model and the deter-
was also reported for Urmu mulberry concentrate (Kara and
mination coefficients (R2) are shown in Table 4. Fitting accuracy was
Erçelebi, 2013).
evaluated through the analysis of R2 coefficients and the regression
The significant (p < 0.05) decrease in hunter L* value indicates
plot of experimental versus predicted anthocyanin extraction yield
that the extracts became darker at higher temperatures. The
by the model. The Peleg model could adequately predict the effect
treated extracts were 26.31e57.47% darker compared to the control.
of thermal pre-treatment on anthocyanin content of extracts, with
The C* value shows the degree of color intensity. Interestingly,
R2 coefficients ranging from 0.951 to 0.986. The corresponding
non-significant (p > 0.05) changes in C* values was observed, which
response values obtained from the experimental data and those
suggests anthocyanin extracts kept the color vividness during heat
predicted by the Peleg model were close together, with a good
treatment.
correlation coefficient of R2 ¼ 0.964 (Fig. 4).
DE increased during blanching. Different reactions may simul-
During blanching of mangosteen pericarp the positive effect of
taneously contribute to color changes such as thermal degradation
blanching on extraction yield of anthocyanins may have covered
of pigments, oxidation of ascorbic acid, enzymatic browning, and
the possible loss of anthocyanins due to degradation and/or
non-enzymatic browning (Ling et al., 2014). Besides, heating the
leaching. The results of extraction study are beneficial for devel-
anthocyanin solution moves the equilibrium between different
oping extraction processes for mangosteen pericarp.
anthocyanin forms toward the formation of chalcone and caused a
decrease in the colored form of anthocyanin and loss of redness
3.4. Impact of thermal treatment on color properties (Patras et al., 2010). Chalcone is an unstable form and further de-
grades to brown products (Tang et al., 2014).
Color is an important quality attribute of food. Knowledge on The constant rates of color loss and corresponding statistical
color kinetic parameters allows us to predict color changes and to parameters are presented in Table 5. Effect of blanching on the vi-
improve the product appearance through suitable selection of sual red color (a* value), C* and DE was modeled using zero-order
processing conditions. Color properties at all tested conditions model, while first-order model well described the changes of L*
were determined, however, only initial and final values are re- value. Comparing the Ea values of color parameters (Table 6), a* was
ported in Table 5. the most susceptible Hunter color value to thermal treatment. The

Table 4
Peleg's constants and correlation coefficient (R2) for anthocyanin extraction after blanching at different temperatures.

Temperature (oC) K1 (min. 100 g/mg) K2 (100 g/mg) Extracted anthocyanin (mg/100 g)a R2

Observed Predicted

60 3.76 # 10'2 ± 1.37 2.41 # 10'2 ± 0.18 168.94 170.14 0.979


70 3.23 # 10'2 ± 1.41 1.59 # 10'2 ± 0.18 183.44 187.21 0.951
80 2.34 # 10'2 ± 0.97 1.88 # 10'2 ± 0.12 180.14 181.61 0.983
90 1.19 # 10'2 ± 1.43 2.05 # 10'2 ± 0.18 178.16 179.95 0.986
100 8.3 # 10'3 ± 1.3 1.87 # 10'2 ± 0.13 182.06 184.99 0.981
a
The reported results of extracted anthocyanin after 12 min of blanching.
18 M. Ziabakhsh Deylami et al. / Journal of Food Engineering 178 (2016) 12e19

Table 5
Kinetic parameters for a*, L*, C* and total color changes (DE) of mangosteen pericarp extract estimated with proposed kinetic models at different temperatures.

Non-blanched After 12 min of blanching

60 ! C 70 ! C 80 ! C 90 ! C 100 ! C
Aa Bb Bc Bd Be
*
L value 43.63 ± 0.35 33.46 ± 2.01 30.78 ± 1.38 28.66 ± 0.27 27.42 ± 0.95 25.42 ± 0.54Bf
k (#10'2 min'1) 2.2 ± 0.1 2.8 ± 0.2 3.3 ± 0.2 3.9 ± 0.3 4.7 ± 0.2
R2 0.985 0.973 0.989 0.993 0.969
a* value 61.97 ± 0.44Aa 60.04 ± 2.03Bb 54.26 ± 3.85Bc 56.07 ± 3.21Bd 51.92 ± 2.73Be 47.97 ± 4.80Bf
k (#10'1 min'1) 2.27 ± 0.12 4.62 ± 0.44 6.58 ± 0.18 9.21 ± 0.24 11.33 ± 0.68
R2 0.986 0.956 0.996 0.996 0.979
C* value 82.21 ± 0.85Aa 82.15 ± 0.92Aa 82.09 ± 0.71Aa 82.19 ± 1.01Aa 82.00 ± 0.61Aa 81.93 ± 0.84Aa
k (#10'3 min'1) 5.210 ± 0.219 9.756 ± 0.552 13.434 ± 0.387 17.205 ± 0.322 20.918 ± 1.039
R2 0.991 0.984 0.986 0.978 0.988
DE* value 0Aa 10.52 ± 0.46Bb 14.23 ± 0.52Bc 17.71 ± 0.22Bd 21.71 ± 0.49Be 25.12 ± 0.68Bf
k (min'1) 1.082 ± 0.050 1.227 ± 0.113 1.848 ± 0.100 2.219 ± 0.111 2.410 ± 0.152
R2 0.987 0.951 0.983 0.985 0.977

The capital letters indicate statistically significant differences (p < 0.05) between blanched and non-blanched sample. The small letters show statistically significant differences
(p < 0.05) between the mean values for samples blanched at different temperatures.

Table 6
Activation energy values for PPO inactivation, anthocyanin degradation and color changes due to thermal treatment.

PPO inactivation Anthocyanin degradation Color changes

L* a* C* DE
Ea (kJ/mol) 43.11 ± 2.22 57.66 ± 3.66 19.09 ± 0.44 32.22 ± 2.62 27.02 ± 2.96 18.86 ± 3.37
R2 0.983 0.955 0.998 0.968 0.948 0.882

higher degradation of red color could be due to anthocyanin Batista, K.A., Batista, G.G.L.A., Alves, G.L., Fernandes, K.F., 2014. Extraction, partial
purification and characterization of polyphenol oxidase from Solanum lyco-
degradation of the extracts. The perception of color in treated
carpum fruits. J. Mol. Catal. B Enzym. 102, 211e217. http://dx.doi.org/10.1016/
samples is the result of change in L*, a* and b* (Kara and Erçelebi, j.molcatb.2014.02.017.
2013). Considering these values independently does not offer in- Boussetta, N., Lanoiselle !, J.L., Bedel-Cloutour, C., Vorobiev, E., 2009. Extraction of
formation about intensity and tone of the color. Therefore, it is soluble matter from grape pomace by high voltage electrical discharges for
polyphenol recovery: effect of sulphur dioxide and thermal treatments. J. Food
suggested to use DE along with other quality attributes in order to Eng. 95, 192e198. http://dx.doi.org/10.1016/j.jfoodeng.2009.04.030.
judge the preferred timeetemperature combination for blanching Bridle, P., Timberlake, C.F., 1997. Anthocyanins as natural food coloursdselected
of mangosteen pericarp. aspects. Food Chem. 58, 103e109.
Brownmiller, C., Howard, L.R., Prior, R.L., Hager, A., 2008. Processing and storage
effects on monomeric anthocyanins, percent polymeric color, and antioxidant
4. Conclusions capacity of processed blueberry products. J. Food Sci. 73, H72eH79. http://
dx.doi.org/10.1111/j.1750-3841.2008.00761.x.
Buci!
c-Koji! c, A., Planini!
c, M., Tomas, S., Bili!
c, M., Veli!
c, D., 2007. Study of solideliquid
PPO inhibition as the first step for stabilization of anthocyanins extraction kinetics of total polyphenols from grape seeds. J. Food Eng. 81,
of mangosteen pericarp was studied. The results showed quite high 236e242. http://dx.doi.org/10.1016/j.jfoodeng.2006.10.027.
Cheng, X., Zhang, M., Adhikari, B., 2013. The inactivation kinetics of polyphenol
thermal stability of the enzyme. Considering the thermal degra- oxidase in mushroom (Agaricus bisporus) during thermal and thermosonic
dation of anthocyanins prolonged blanching for complete enzyme treatments. Ultrason. Sonochem 20, 674e679. http://dx.doi.org/10.1016/
inhibition was not applicable. Therefore, blanching as a single j.ultsonch.2012.09.012.
Chiste!, R.C., Lopes, A.S., De Faria, L.J.G., 2010. Thermal and light degradation kinetics
pretreatment method was not sufficient to inhibit enzyme activity. of anthocyanin extracts from mangosteen peel (Garcinia mangostana L.). Int. J.
If mangosteen pericarp will be processed into paste or powder or Food Sci. Technol. 45, 1902e1908. http://dx.doi.org/10.1111/j.1365-
will be stored for a while, it is suggested to use blanching in com- 2621.2010.02351.x.
Cipriano, P.D.A., Ekici, L., Barnes, R.C., Gomes, C., Talcott, S.T., de Aguiar Cipriano, P.,
bined with another inhibition method, such as chemical treatment, Ekici, L., Barnes, R.C., Gomes, C., Talcott, S.T., 2015. Pre-heating and polyphenol
to achieve complete PPO inhibition. Blanching enhanced anthocy- oxidase inhibition impact on extraction of purple sweet potato anthocyanins.
anin extraction efficiency; therefore it is highly recommended as a Food Chem. 180, 227e234. http://dx.doi.org/10.1016/j.foodchem.2015.02.020.
Cisse, M., Vaillant, F., Acosta, O., Dhuique-Mayer, C., Dornier, M., 2009. Thermal
processing step prior to anthocyanin extraction.
degradation kinetics of anthocyanins from blood orange, blackberry, and roselle
using the arrhenius, eyring, and ball models. J. Agric. Food Chem. 57,
6285e6291. http://dx.doi.org/10.1021/jf900836b.
References Corradini, M.G., Peleg, M., 2004. A model of non-isothermal degradation of nutri-
ents, pigments and enzymes. J. Sci. Food Agric. 84, 217e226. http://dx.doi.org/
Agüero, M.V., Ansorena, M.R., Roura, S.I., del Valle, C.E., 2008. Thermal inactivation 10.1002/jsfa.1647.
of peroxidase during blanching of butternut squash. LWT Food Sci. Technol. 41, Dembitsky, V.M., Poovarodom, S., Leontowicz, H., Leontowicz, M., Vearasilp, S.,
401e407. http://dx.doi.org/10.1016/j.lwt.2007.03.029. Trakhtenberg, S., Gorinstein, S., 2011. The multiple nutrition properties of some
Ahmed, J., Shivhare, U.S., Raghavan, G.S.V., 2004. Thermal degradation kinetics of exotic fruits: biological activity and active metabolites. Food Res. Int. 44,
anthocyanin and visual colour of plum puree. Eur. Food Res. Technol. 218, 1671e1701. http://dx.doi.org/10.1016/j.foodres.2011.03.003.
525e528. http://dx.doi.org/10.1007/s00217-004-0906-5. Falguera, V., S! anchez-Rian ~ o, A.M., Quintero-Cero !n, J.P., Rivera-Barrero, C.A.,
Altunkaya, A., Go €kmen, V., 2012. Partial purification and characterization of poly- Me !ndez-Arteaga, J.J., Ibarz, A., Banat, B.F., Jumah, R., Hammad, S., 2012. Char-
phenoloxidase from durum wheat ( Triticum durum L.). J. Cereal Sci. 55, acterization of polyphenol oxidase activity in juices from 12 underutilized
300e304. http://dx.doi.org/10.1016/j.jcs.2011.12.013. tropical fruits with high agroindustrial potential. Food Bioprocess Technol. 2,
Anthon, G.E., Barrett, D.M., 2002. Kinetic parameters for the thermal inactivation of 403e407. http://dx.doi.org/10.1007/s11947-011-0521-y.
quality-related enzymes in carrots and potatoes. J. Agric. Food Chem. 50, Fang, Z., Zhang, M., Sun, Y., Sun, J., 2007. Polyphenol oxidase from bayberry (Myrica
4119e4125. rubra Sieb. et Zucc.) and its role in anthocyanin degradation. Food Chem. 103,
Arnnok, P., Ruangviriyachai, C., Mahachai, R., Techawongstien, S., Chanthai, S., 2010. 268e273. http://dx.doi.org/10.1016/j.foodchem.2006.07.044.
Optimization and determination of polyphenol oxidase and peroxidase activ- Galva!n D'Alessandro, L., Dimitrov, K., Vauchel, P., Nikov, I., 2014. Kinetics of
ities in hot pepper (Capsicum annuum L.) pericarb. Int. Food Res. J. 17, 385e392.
M. Ziabakhsh Deylami et al. / Journal of Food Engineering 178 (2016) 12e19 19

ultrasound assisted extraction of anthocyanins from Aronia melanocarpa (black dx.doi.org/10.1016/j.postharvbio.2008.08.003.


chokeberry) wastes. Chem. Eng. Res. Des. 92, 1818e1826. http://dx.doi.org/ Patras, A., Brunton, N.P., O'Donnell, C., Tiwari, B.K., 2010. Effect of thermal pro-
10.1016/j.cherd.2013.11.020. cessing on anthocyanin stability in foods; mechanisms and kinetics of degra-
Ganjloo, A., Rahman, R.A., Osman, A., Bakar, J., Bimakr, M., 2011. Kinetics of crude dation. Trends Food Sci. Technol. 21, 3e11. http://dx.doi.org/10.1016/
peroxidase inactivation and color changes of thermally treated seedless guava j.tifs.2009.07.004.
(Psidium guajava L.). Food bioprocess Technol. 4, 1442e1449. http://dx.doi.org/ Peterson, G.L., 1977. A simplification of the protein assay method of Lowry et al.
10.1007/s11947-009-0245-4. which is more generally applicable. Anal. Biochem. 83, 346e356. http://
Garcia-Palazon, A., Suthanthangjai, W., Kajda, P., Zabetakis, I., 2004. The effects of dx.doi.org/10.1016/0003-2697(77)90043-4.
high hydrostatic pressure on b-glucosidase, peroxidase and polyphenoloxidase Rossi, M., Giussani, E., Morelli, R., Lo Scalzo, R., Nani, R.C., Torreggiani, D., 2003.
in red raspberry (Rubus idaeus) and strawberry (Fragaria#ananassa). Food Effect of fruit blanching on phenolics and radical scavenging activity of high-
Chem. 88, 7e10. http://dx.doi.org/10.1016/j.foodchem.2004.01.019. bush blueberry juice. Food Res. Int. 36, 999e1005.
Giusti, M.M., Wrolstad, R.E., 2001. Characterization and measurement of anthocy- Ruenroengklin, N., Sun, J., Shi, J., Xue, S.J., Jiang, Y., 2009. Role of endogenous and
anins by UV-visible spectroscopy. Curr. Protoc. Food Anal. Chem. exogenous phenolics in litchi anthocyanin degradation caused by polyphenol
Gonçalves, E.M., Pinheiro, J., Abreu, M., Branda ~o, T.R.S., Silva, C.L.M., 2010. Carrot oxidase. Food Chem. 115, 1253e1256. http://dx.doi.org/10.1016/
(Daucus carota L) peroxidase inactivation, phenolic content and physical j.foodchem.2009.01.040.
changes kinetics due to blanching. J. Food Eng. 97, 574e581. Shivhare, U.S., Gupta, M., Basu, S., Raghavan, G.S.V., 2009. Optimization of blanching
He, J., Giusti, M.M., 2010. Anthocyanins: natural colorants with health-promoting process for carrots. J. Food Process Eng. 32, 587e605. http://dx.doi.org/10.1111/
properties. Annu. Rev. Food Sci. Technol. 2010 (1), 163e187. http://dx.doi.org/ j.1745-4530.2007.00234.x.
10.1146/annurev.food.080708.100754 new. Stamatopoulos, K., Katsoyannos, E., Chatzilazarou, A., Konteles, S.J., 2012.
Kader, F., Rovel, B., Girardin, M., Metche, M., 1997. Mechanism of browning in fresh Improvement of oleuropein extractability by optimising steam blanching pro-
highbush blueberry fruit (Vaccinium corymbosum L). Role of blueberry poly- cess as pre-treatment of olive leaf extraction via response surface methodology.
phenol oxidase, chlorogenic acid and anthocyanins. J. Sci. Food Agric. 74, 31e34. Food Chem. 133, 344e351.
Kara, Ş., Erçelebi, E.A., 2013. Thermal degradation kinetics of anthocyanins and vi- Sui, X., Dong, X., Zhou, W., 2014. Combined effect of pH and high temperature on
sual colour of Urmu mulberry ( Morus nigra L.). J. Food Eng. 116, 541e547. the stability and antioxidant capacity of two anthocyanins in aqueous solution.
http://dx.doi.org/10.1016/j.jfoodeng.2012.12.030. Food Chem. http://dx.doi.org/10.1016/j.foodchem.2014.04.075.

Kırca, A., Ozkan, M., Cemerog #lu, B., 2007. Effects of temperature, solid content and Tan, T.-C., Cheng, L.-H., Bhat, R., Rusul, G., Easa, A.M., 2014. Composition, physico-
pH on the stability of black carrot anthocyanins. Food Chem. 101, 212e218. chemical properties and thermal inactivation kinetics of polyphenol oxidase
http://dx.doi.org/10.1016/j.foodchem.2006.01.019. and peroxidase from coconut (Cocos nucifera) water obtained from immature,
Klabunde, T., Eicken, C., Sacchettini, J.C., Krebs, B., 1998. Crystal structure of a plant mature and overly-mature coconut. Food Chem. 142, 121e128. http://
catechol oxidase containing a dicopper center. Nat. Struct. Mol. Biol. 5, dx.doi.org/10.1016/j.foodchem.2013.07.040.
1084e1090. http://dx.doi.org/10.1038/4193. Tang, K., Li, Y., Han, Y., Han, F., Li, J., Nie, Y., Xu, Y., 2014. Studies on the preparative
Liazid, a., Guerrero, R.F., Cantos, E., Palma, M., Barroso, C.G., 2011. Microwave isolation and stability of seven main anthocyanins from Yan 73 grape. J. Sci.
assisted extraction of anthocyanins from grape skins. Food Chem. 124, Food Agric. 94, 2472e2481. http://dx.doi.org/10.1002/jsfa.6582.
1238e1243. http://dx.doi.org/10.1016/j.foodchem.2010.07.053. Terefe, N.S., Yang, Y.H., Knoerzer, K., Buckow, R., Versteeg, C., 2010. High pressure
Lin, L., Lei, F., Sun, D.-W., Dong, Y., Yang, B., Zhao, M., 2012. Thermal inactivation and thermal inactivation kinetics of polyphenol oxidase and peroxidase in
kinetics of Rabdosia serra (Maxim.) Hara leaf peroxidase and polyphenol oxi- strawberry puree. Innov. Food Sci. Emerg. Technol. 11, 52e60. http://dx.doi.org/
dase and comparative evaluation of drying methods on leaf phenolic profile and 10.1016/j.ifset.2009.08.009.
bioactivities. Food Chem. 134, 2021e2029. http://dx.doi.org/10.1016/ Ünal, M., 2007. Properties of polyphenol oxidase from Anamur banana (Musa
j.foodchem.2012.04.008. cavendishii). Food Chem. 100, 909e913. http://dx.doi.org/10.1016/
Ling, B., Tang, J., Kong, F., Mitcham, E.J., Wang, S., 2014. Kinetics of food quality j.foodchem.2005.10.048.
changes during thermal processing: a review. Food Bioprocess Technol. http:// Van Boekel, M.A.J.S., 2008. Kinetic modeling of food quality: a critical review.
dx.doi.org/10.1007/s11947-014-1398-3. Compr. Rev. Food Sci. Food Saf. 7, 144e158. http://dx.doi.org/10.1111/j.1541-
Navarro, J.L.J., T!arrega, A., Sentandreu, M.M.A., Sentandreu, E., 2014. Partial purifi- 4337.2007.00036.x.
cation and characterization of polyphenol oxidase from persimmon. Food Volden, J., Borge, G.I.A., Bengtsson, G.B., Hansen, M., Thygesen, I.E., Wicklund, T.,
Chem. 157, 283e289. http://dx.doi.org/10.1016/j.foodchem.2014.02.063. 2008. Effect of thermal treatment on glucosinolates and antioxidant-related
Nayak, B., Berrios, J.D.J., Powers, J.R., Tang, J., 2011. Thermal degradation of antho- parameters in red cabbage (Brassica oleracea L. ssp. capitata f. rubra). Food
cyanins from purple potato (cv. Purple Majesty) and impact on antioxidant Chem. 109, 595e605.
capacity. J. Agric. Food Chem. 59, 11040e11049. http://dx.doi.org/10.1021/ Yang, C.P., Fujita, S., Kohno, K., Kusubayashi, A., Ashrafuzzaman, M., Hayashi, N.,
jf201923a. 2001. Partial purification and characterization of polyphenol oxidase from ba-
Odriozola-Serrano, I., Soliva-Fortuny, R., Martín-Belloso, O., 2009. Influence of nana (Musa sapientum L.) peel. J. Agric. Food Chem. 49, 1446e1449.
storage temperature on the kinetics of the changes in anthocyanins, vitamin C, Zarena, a., Udaya Sankar, K., 2012. Isolation and identification of pelargonidin 3-
and antioxidant capacity in fresh-cut strawberries stored under high-oxygen glucoside in mangosteen pericarp. Food Chem. 130, 665e670. http://dx.doi.org/
atmospheres. J. Food Sci. 74, 184e191. http://dx.doi.org/10.1111/j.1750- 10.1016/j.foodchem.2011.07.106.
3841.2009.01075.x. Ziabakhsh Deylami, M., Russly, A.R., Tan, C.P., Bakar, J., Olusegun, L., 2014. Ther-
Palapol, Y., Ketsa, S., Stevenson, D., Cooney, J.M., Allan, A.C., Ferguson, I.B., 2009. modynamics and kinetics of thermal inactivation of peroxidase from mango-
Colour development and quality of mangosteen ( Garcinia mangostana L.) fruit steen (Garcinia mangostana l.) pericarp. J. Eng. Sci. Technol. 9, 374e383.
during ripening and after harvest. Postharvest Biol. 51, 349e353. http://

You might also like