Você está na página 1de 19

PDFlib PLOP: PDF Linearization, Optimization, Protection

Page inserted by evaluation version


www.pdflib.com sales@pdflib.com

Blackwell Science, LtdOxford, UKJFPPJournal of Food Processing and Preservation0145-8892Copyright 2005 by Food & Nutrition Press, Inc., Trumbull, Connecticut.2914562Original Article STORAGE,

SHELF LIFE AND COLOR OF

MANGO POWDERS. JAYA and H. DAS

ACCELERATED STORAGE, SHELF LIFE AND COLOR OF


MANGO POWDER
S. JAYA and H. DAS1
Department of Agricultural and Food Engineering
Indian Institute of Technology, Kharagpur
India 721 302
Accepted for Publication March 1, 2005

ABSTRACT
Vacuum-dried mango powder was produced from mango pulp through
the addition of glycerol monostearate and tricalcium phosphate at 0.015 kg
each per kg mango solids and maltodextrin at 0.62 kg per kg dry mango
solids. The mango powder was packed in aluminum foil-laminated pouches
and stored in an accelerated storage environment maintained at 90% relative
humidity (RH) and 38 2C. The sticky-point moisture content at 38 2C was
considered as the maximum moisture content to which the mango powder
would remain stable. The shelf life of the powder predicted from this consideration and the Guggenheim-Anderson-de Boer (GAB) model for the water
activity moisture content relationship was 114.68 days, whereas the actual
shelf life was 105 days. The color change of the powder during storage
followed first-order reaction kinetics with a rate constant of 0.038 per day.

INTRODUCTION
The drying of fruit pulp to powder is difficult mainly because of low
molecular weight sugars such as fructose, glucose, sucrose and acids such as
citric acid present in the pulp. These sugars and acids exhibit low glass
transition temperatures (Roos and Karel 1991; Roos 1995), and because of
their low molecular weight, the molecular mobility of the materials is high
when the temperature is just above the glass transition temperature. Because
of the short duration of time that is normally available for drying, dried sugars
and acids attain the amorphous state (Masters 1985).
The glass transition temperature of a food is the temperature at which an
amorphous food polymer is transformed into the viscous liquid or the rubbery
1

Corresponding author.
hd@agfe.iitkgp.ernet.in

TEL:

91-03222-283156;

FAX:

91-03222-282244;

Journal of Food Processing and Preservation 29 (2005) 4562. All Rights Reserved.
Copyright 2005, Blackwell Publishing

EMAIL:

45

46

S. JAYA and H. DAS

state from the solid amorphous state. A dramatic change in molecular mobility
and physical properties takes place at the glass transition temperature (Slade
and Levine 1991).
The temperature at which amorphous substances exhibit stickiness is
about 1020C higher than the glass transition temperature (Bhandari et al.
1993, 1997; Roos 1995). Because low glass transition temperatures increase
the stickiness, it is important to increase the glass transition temperature of
fruit powders by adding high molecular weight materials such as maltodextrin
and lactose (Roos and Karel 1991; Slade and Levine 1994; Roos 1995; Roos
et al. 1996; Bhandari et al. 1997).
The cause of stickiness is plasticization of low molecular weight sugar,
present because of water sorption and subsequent interparticle fusion. The
sticky-point temperature is the temperature at which caking is instantaneous
with slow stirring. A sticky-point temperature apparatus developed by Lazar
et al. (1956) was used by several others (Downton et al. 1982; Wallack and
King 1988; Buhler and Liedy 1989; Jaya et al. 2002). The test assesses the
influence of temperature and moisture content on powder cohesion and is
applicable to nearly dry powders, which form immobile liquid bridges of
water present in the particles (Papadakis and Babu 1992). When the stickypoint temperature drops below room temperature, the food will develop stickiness (Roos 1995).
A relationship between moisture content and sticky-point temperature
can be developed by using the apparatus developed by Lazar et al. (1956). By
determining the sticky-point moisture content of foods at selected storage
temperatures and by selecting the sticky-point moisture content as the critical
moisture content, the quality storage life of food in the given storage environments (Jaya et al. 2002) can be predicted. Also, the maximum storage temperature for a food can be obtained from the sticky-point temperature and
moisture content relationship.
Powdered dehydrated products of fruit juice, soup, custard etc. require
protection against ingress of moisture and oxygen and the loss of volatile flavorings and color. Powders are usually packed in heat-sealable laminates containing a layer or layers of aluminum, such as aluminum foil-laminated
polyethylene. Accelerated storage involving high humidity and temperature
such as 90% relative humidity (RH) and 38 2C can be used for developing
moisture ingress and storage time relationships quickly (Potter 1978). Using
this relationship and the critical moisture content, the shelf life of a dry food
can be predicted. A model for shelf life prediction that is based on sticky-point
moisture content, permeability of the packaging material and storage temperature will be useful for identifying the quality storage life of fruit juice powder.
Because the removal of moisture takes place in the absence of oxygen
in vacuum drying, oxidative degradation, e.g., browning, is very small in the

STORAGE, SHELF LIFE AND COLOR OF MANGO POWDER

47

final product. Temperatures for vacuum drying are normally kept below 75C
(Copley et al. 1956; Pap 1995). During vacuum drying, fruit juice/pulp droplets expand as the air and water vapor in the juice/pulp develop a frothy, puffed
structure. The expanded structure provides large area to volume ratios, promoting rapid heat and mass transfer and consequently, high rates of drying.
Being a surface active agent, the glycerol monostearate present in food
reduces the surface tension in the food and promotes the development of
puffed structures when a vacuum is applied. The formation of the puffed
structure hastens the vacuum-drying process (Morgan et al. 1959). The production of a free-flowing fruit powder is enhanced by the incorporation of
food-grade anticaking agents such as tricalcium phosphate added to fruit pulp.
Tricalcium phosphate binds water by competing with other materials present
in food (James 1971; Peleg and Hollenbech 1984).
Fruit powder can be produced by adding maltodextrin, glycerol
monostearate and tricalcium phosphate to fruit pulp and by drying the pulp
in a vacuum dryer. Because of the ingress of moisture through the packaging
materials, the powder exhibits stickiness and color change during storage. The
objectives of this study were: (1) to develop a prediction model for the shelf
life of mango powder based on sticky-point temperature, moisture content and
permeability of packaging material and (2) to observe the color changes of
mango powder using accelerated storage conditions.

MATERIALS AND METHODS


Analysis of Mango Pulp
A survey on the mango varieties used by processors for the production
of mango drinks revealed that Totapuri mango cultivar is preferred over other
cultivars. Canned Totapuri mango pulp (Tulip Products, Madhyamgram, West
Bengal) was used for the production of mango powder. The mango pulp was
analyzed for reducing sugars, total sugars, sucrose, fructose, glucose, acidity
(as anhydrous citric acid) and moisture, following methods prescribed by
Ranganna (1987) and Sadasivam and Manickam (1996). Table 1 presents the
composition of mango pulp.
Drying of Mango Pulp
On the basis of the quality criteria, viz., hygroscopicity, degree of caking,
dispersibility, flowability and sticky-point temperature of vacuum-dried
mango pulp, optimum amounts of maltodextrin, glycerol monostearate and
tricalcium phosphate were identified as 0.62, 0.015 and 0.015 kg/kg dry
mango solids, respectively. The mango pulp was mixed with the ingredients

48

S. JAYA and H. DAS

TABLE 1.
MEAN COMPOSITION OF TOTAPURI CULTIVAR OF MANGO PULP
Constituents

Amount (kg per kg pulp)

Moisture content
Total solids
Sucrose
Fructose
Glucose
Total sugar
Acidity as anhydrous citric acid

0.800.85
0.150.20
0.06
0.025
0.046
0.130.14
0.0031

using a planetary mixer operated at 75-m/min peripheral speed for 1520 min.
The mix was then heated to 70C and spread to 3-mm thickness (Pap 1995)
on a Teflon-coated tray inserted in the vacuum-drying chamber. The thickness
of the mango pulp mix was fixed from weight of pulp, tray surface area and
density of the pulp. The tray surface area and the density of the pulp mixed
with the ingredients were 0.0755 m2 and 1115 5 kg/m3, respectively.
The amount of mango pulp used in each experiment was 0.2 kg. The mix
was dried in the dryer at 70 2C (Copley et al. 1956; Pap 1995) and 710- to
750-mmHg vacuum. A water ring-type vacuum pump was fitted to the dryer
to remove water vapor from the pulp during drying. Because the water vapor
was continuously removed from the vacuum dryer, the rate of drying was
greater than the rate of drying in a vacuum oven. During drying, the tray was
removed at 15-min intervals and weighed. The moisture contents of the dry
mango pulp were calculated from the weights of the mango pulp before and
after drying and the initial moisture content of the pulp. The total drying time
was approximately 1.75 0.25 h. After drying to a moisture content less than
about 0.001 kg/kg dried mango pulp, the dried mango pulp was removed from
the dryer and cooled in an ambient atmosphere until the mango pulp reached
the ambient temperature. The cooling time was approximately 23 min. The
dried mango pulp was lathery but not sticky.
Conditioning and Grinding of Dried Mango Pulp
The dried mango pulp was conditioned in a low RH (34%) environment
for about 3045 min to attain the desired brittleness. The low RH environment
was created by blowing atmospheric air through a saturated solution of sodium
hydroxide and then through a bed of silica gel. The grinding of the conditioned
mango pulp was carried out in a hammer mill, in a room maintained at 20 2C
and 4050% RH by using a room air conditioner. The mass mean diameter

STORAGE, SHELF LIFE AND COLOR OF MANGO POWDER

49

of the mango powder obtained from the mill was 0.246 mm. The mango
powder was collected in aluminum foil-laminated polyethylene pouches. The
thickness and the unit weight of the pouch material were 90 micron and
0.099 kg/m2, respectively. The weight of the powder obtained from a single
experiment varied between 15 and 60 g.
Measures of Mango Powder Quality
Hygroscopicity. Hygroscopicity is the ability of powder to absorb moisture from high RH environments (Haugaard et al. 1978; Slade and Levine
1991). Hygroscopicity was measured as the equilibrium moisture content of
the mango powder exposed to air at 79.5% RH (Haugaard et al. 1978; Pisecky
1985).
Degree of Caking. The degree of caking is expressed as the percentage
of powder retained on a sieve of defined size after redrying and sieving the
cake formed in the hygroscopicity test (Haugaard et al. 1978; Pisecky 1985).
The tendency to caking was expressed as the percentage of powder retained
on a sieve of 500-micron size after redrying and sieving the cake formed in
the hygroscopicity test.
Dispersibility. Dispersibility is the ability of powder to get wet without
the formation of lumps in water. The International Dairy Federation method
for dispersibility measurement specifies a stirring procedure followed by pouring the solution through a 210-mm sieve (Pisecky 1985). Weighing the dry
matter present in reconstituted mango juice after passing through the designated sieve and expressing the weight as the percentage of the dry mango
powder used in the reconstitution is the estimated dispersibility.
Flowability. Flowability is an important but complex property controlled by particle size distribution, angle of repose, and moisture content of
powders. Flowability was determined as the time in seconds necessary for a
selected volume of powder (moisture content < 5% db) to pass through slots
of specific size from a rotating drum (Pisecky 1985).
Sticky-point Temperature. The sticky-point temperature, Ts (C), is the
temperature at which the particles that make up a powder start to stick to each
other or to the wall of the container, in which it is enclosed, as the powder in
the container is subjected to increasing temperatures. The apparatus used for
the measurement of sticky point is presented in Fig. 1. The entire assembly,
consisting of sample holder (1), powder (2), sealing liquid holder (3), sealing
tube (5), stirring rod (7) and propeller (9), was immersed into a constant

50

S. JAYA and H. DAS

6
7

3
1
4

11

10

Sealing liquid
(Mercury)

FIG. 1. SCHEMATIC DIAGRAM OF STICKY-POINT TEMPERATURE MEASUREMENT


APPARATUS
(1) Sample holder; (2) powder sample; (3) sealing liquid holder; (4) glass tube; (5) sealing tube; (6)
teflon ring stopper; (7) stainless steel stirring rod; (8) rubber cork; (9) propeller; (10) thermocouple
wire; (11) adhesive seal.

temperature water bath. The temperature of the bath was slowly increased,
and the powder was intermittently stirred by hand by rotating the sealing tube
(5). At a selected temperature, the force required to rotate the propeller inside
the powder increased dramatically. The temperature at which the dramatic
increase in force occurred was usually distinct and referred to as the sticky-

STORAGE, SHELF LIFE AND COLOR OF MANGO POWDER

51

point temperature of the powder. The sticky-point temperature was determined


by using a thermocouple (10) and a digital temperature indicator. The stickypoint temperatures of mango powder at selected moisture contents were determined with this procedure and apparatus.
Color of Powder. A selected quantity of powder composed of a known
amount of total solids (0.01 kg per kg of reconstituted juice) was reconstituted
with water at 30C. The L, a and b values of the reconstituted powder were
determined with a HunterLab Colorimeter (HunterLab D25, Hunter Associates Laboratory, Reston, VA). The total color differences (D E) in the mango
powder between the initial day and the subsequent q days of storage were
calculated using Eq. (1):
DE =

( Lo - Lp )2 + (ao - ap )2 + (bo - bp )2

(1)

where Lo, ao and bo, are the L, a and b values of the mango pulp initially, and
Lp, ap and bp are the corresponding values of the reconstituted powder after q
days of storage.
Water Activity and Moisture Content. The water activity (aw) of the
mango powder was determined using an Aqua lab (Model CX2, Decagon
Devices, Pullman, WA) water activity meter. The moisture content of the
mango powder was determined by using a Metler moisture analyzer (Metler
Toledo AG, Greifensee, Switzerland). The mango powder was assayed at a
temperature of 100C in the infrared drying chamber.
Optimization of Ingredients
Because of the absence of any specified range of values for hygroscopicity, degree of caking, dispersibility, flowability and sticky-point temperature
of fruit juice powders, these values were determined with commercially available tomato soup and instant coffee powders. The mango powder was considered to be of desirable quality if its quality values were within the range of
the quality values of the tomato and coffee powders. The experiments were
conducted with selected concentrations (kg/kg mango solids) of maltodextrin:
0.25, 0.35, 0.45, 0.55 and 0.65; glycerol monostearate: 0.01. 0.0125, 0.015,
0.0175 and 0.02 and tricalcium phosphate: 0.01, 0.0125, 0.015, 0.0175 and
0.02. The optimization of the ingredients was calculated by developing second-order regression equations between the amount of the ingredients added
and the assessed quality values and by solving the equations using Microsoft
Excel solver. The quality values of the mango powder produced with the
optimum amount of the selected ingredients, tomato soup and instant coffee
powders are presented in Table 2.

52

S. JAYA and H. DAS

TABLE 2.
PROPERTIES OF MANGO, INSTANT COFFEE AND TOMATO SOUP POWDERS
Property

Instant coffee
powder

Tomato soup
powder

Mango
powder*

Hygroscopicity (%)
Degree of caking (%)
Dispersibility (%)
Flowability at 35% (db) moisture (s)
Sticky point temperature at 35% (db)
moisture (C)

9.0910.32
31.2734.48
99.9599.98
1518
46 1

4.216.64
8.9610.23
68.1970.21
3642
>70

7.2
18.38
68.67
21.7
52.8

* Prepared from mango pulp added with maltodextrin, glycerol monostearate and tricalcium phosphate at rates of 0.62, 0.015 and 0.015 kg/kg dry mango solid.

Measurement of Permeability of Packaging Material


To determine water vapor permeability of the packaging material, 5 g of
dehydrated silica gel was packed separately in two aluminum foil-laminated
polyethylene (14 10 cm) pouches. The two pouches were placed in an environment maintained at 90% RH and 38 2C established with saturated potassium nitrate solution in an incubator. The temperature inside the incubator
was maintained by using a contact thermometer-type temperature sensorcontroller. The weight of the pouches containing the silica gel was determined
at 24-h intervals for 7 days, and the mean weight gain by the silica gel was
calculated for each day (Labuza 1984). The water vapor permeability, K (kg/
m2/day/Pa), of the packaging material was calculated using Eq. (2):
K=

dw dq p
Ap p *

(2)

where dw/dqp is the slope of the linear plot between the time qp (day) of
incubation and cumulative moisture gain; w (kg) is the weight of the silica
gel in the packaging material; Ap (m2) is the surface area of the packaging
material and p* (Pa) is the saturation vapor pressure of water at 38C, the
temperature of the environment chosen for the experiment.
Accelerated Storage of Mango Powder
Twenty grams of mango powder with an initial moisture content
Xi = 0.004 kg water/kg dry solids was packed in 15 15 cm pouches and
placed in an environment maintained at 90% RH and 38 2C. Seven such
pouches were prepared, and after intervals of 15 days, one of the pouches was
removed from the control environment and its contents analyzed for (1)

STORAGE, SHELF LIFE AND COLOR OF MANGO POWDER

53

moisture content X (kg water/kg dry solids); (2) water activity and (3) color
difference.
GAB Model Constants of Mango Powder
The moisture content of the mango powders stored inside the package
increased because of the migration of water vapor from the storage environment through the packaging material. Because the stickiness of mango powder
is a function of its moisture content and temperature, the powder exhibited
stickiness after its moisture content was increased to a critical moisture
content corresponding to the temperature of the storage environment, i.e.,
38 2C. This moisture content is designated as the critical moisture content
Xc (kg water/kg dry solids) of the mango powder stored at 38 2C.
It was assumed that the particles of the powder inside the pouches gained
moisture uniformly. The water activity and moisture content data was fitted
to the Guggenheim-Anderson-de Boer (GAB) (Rockland and Beuchat 1987)
equation expressed as
2
Mo
Mo
- 1 C - 2 +
2+
- 1 C - 4(1 - C)
X
X

aw =
2 k(1 - C)

) ( (

))

0.5

(3)

where aw is water activity; X (kg water per kg of dry solids) is the equilibrium
moisture content; C and k are constants and Mo (kg water per kg of dry solids)
is the monolayer moisture content of the powder. A realistic set of aw - X data
available from the experiment, were used to estimate the best fit values of the
constants C, k and Mo of Eq. (3).
Moisture Gain and Storage Life Prediction
The rate of change of moisture content dX/dq of the powder with storage
time q is expressed as
dX
(4)
= KAp ( Rhp* - aw p*)
Ws
dq
where Ws (kg) is the dry weight of the powder inside the pouch; p* (Pa) is
the saturation vapor pressure of water at the temperature T (C) of storage; Rh
is the relative humidity of the storage environment; K (kg water/m2/day/Pa)
is the permeability of the packaging material; Ap (m2) is the surface area of
the packaging material through which water vapor permeates; aw is the water
activity of the powder at T (C) and X (kg water/kgdrysolids) is the moisture
content of the powder after q days of storage time.
The graphical relationship between the moisture content X (kg water/kg
dry solids) and the storage time q (days) was obtained by solving Eq. (4). The

54

S. JAYA and H. DAS

computed values of moisture content were compared with the experimentally


determined moisture content. The time q (days) required for the moisture
content of the powder to increase from an initial value of Xi (kg water/kg dry
solids) to its critical value Xc (kg water/kg dry solids) was obtained from
Eq. (4). The initial value of the moisture content, Xi, will normally be smaller
than the monolayer value, Mo (kg water per kg dry solids) of the powder and
is obtained from the GAB equation (Eq. 3). As moisture migrates through the
packaging material, the water activity of the powder increases, and the RH of
the environment decreases.
To check the validity of this model, the mean relative percent derivation
modulus, E (%), was calculated by using the relationship reported by Lemon
et al. (1985):
100 N Xa - X
(5)
E=
X
N i =1
where E (%) is the mean relative percentage deviation modulus; N is the total
number of observations; Xa is the actual moisture content of the powder and
X is the predicted value of the moisture content of the powder obtained from
Eq. (4). A value of E less than 10% is considered as a reasonably good fit for
the model to the real experimental values (Lemon et al. 1985).
Kinetics of Color Change During Storage
From the determination of overall color differences, DE, of reconstituted
mango powder during storage in accelerated storage conditions, the rate constant k (1/day) and the order of reaction n (Lau et al. 2000) were estimated
from Eq. (6):
d ( D E* - D E )
(6)
= - kc ( D E * - D E ) n
dq
where DE* is the maximum possible overall color difference of mango powder. The value of DE* was obtained from a graphical plot of the experimental
values of DE and q. Solving Eq. (6) for n = 0 (zero order), n = 1 (first order)
and n = 2 (second order), and considering that DE = 0 at q = 0, the equations
derived are

zero order: DE* - DE = DE* + k0q

(7)

first order: ln(DE* - DE) = ln(DE* + k1q)

(8)

1
1
(9)
=
-k q
( D E* - D E ) D E* 2
where, k0, k1 and k2 are reaction rate constants for the zero-order, first-order
and second-order reactions, respectively. From Eqs. (7), (8) and (9), the values
second order:

STORAGE, SHELF LIFE AND COLOR OF MANGO POWDER

55

of k0, k1 and k2 can be obtained, respectively, from the slope of the linear plot
between (1) (DE* - DE) and q; (2) ln(DE* - DE) and q and (3) 1/(DE* - DE)
and q. The kinetic order of the reaction responsible for the change in overall
color difference of the mango powder will be the kinetic order corresponding
to the maximum correlation coefficient between the observed and the predicted values of (DE* - DE) obtained from Eqs. (7), (8) and (9).
RESULTS AND DISCUSSION
Composition of Mango Pulp
Table 1 presents the composition of Totapuri cultivar mango pulp
selected for the experiment. The quantities of sucrose, glucose, fructose, total
sugars, moisture content, total solids and acidity were within the ranges cited
in CFTRI (1990) for the Totapuri cultivar of mangoes. The quantities of total
sugars present in the mango pulp is high (0.1310.136 kg per kg of pulp
representing 87.390.7% of total solids) (Table 1). The sticky nature of the
dried mango pulp is attributed to the quantity of total sugars in the total solids
(Roos and Karel 1991). To avoid the stickiness, the addition of maltodextrin
to the pulp was necessary (Bhandari et al. 1997).
The optimum amounts (kg/kg mango solids) of ingredients mixed with
mango pulp were: maltodextrin = 0.62, glycerol monostearate = 0.015 and
tricalcium phosphate = 0.015. With this composition of ingredients and dried
mango pulp, the vacuum-dried mango powder exhibited the hygroscopicity,
degree of caking, dispersibility, flowability and sticky-point temperature
within corresponding quality parameters of tomato soup and instant coffee
powders (Table 2).
Permeability of Packaging Material
The cumulative moisture gain with time by silica gel in aluminum
foil-laminated polyethylene pouches at 38 2C and 90% RH is presented
in Fig. 2. The calculated value of slopes of the best-fit straight lines was
dw/dqp = 0.00001 kg/day. By inserting the surface area of the pouch,
Ap = 0.14 0.1 2 = 0.028 m2 and the saturation vapor pressure, p* of water
at 38C = 6980.5 Pa, into Eq. (2), the value of the water vapor permeability, K,
of the aluminum foil-laminated pouch is 5.4 10-8 kg/m2/day/Pa.
Relationship Between Sticky Point Temperature and Moisture Content
Figure 3 presents the variations in sticky-point temperatures of mango
powder produced with 0.62 kg per kg dry mango solids of maltodextrin and

S. JAYA and H. DAS

Cumulative moisture gain by silica gel (kg)

56

0.00005

y = 1E 05x 3E 08
R 2 = 0.9704

0.000045
0.00004
0.000035
0.00003
0.000025
0.00002
0.000015
0.00001
0.000005
0
0

Time (day)

FIG. 2. CUMULATIVE MOISTURE GAIN BY SILICA GEL THROUGH ALUMINUM-FOIL


LAMINATED POLYETHYLENE POUCH WITH TIME OF STORAGE IN CONTROLLED
ENVIRONMENT

0.015 kg each of glycerol monostearate and tricalcium phosphate per kg dry


mango solids at selected moisture contents. The sticky-point temperature of
the dried mango powder decreases with increases in moisture content. At 0%
moisture, the sticky-point temperature is 72C. At 40C, the extrapolated stickypoint moisture content is 0.089 kg water/kg dry solids.
Shelf Life Prediction of Mango Powder
The GAB model constants were obtained from experimental water
activities and moisture contents (Table 3) as: C = 8.54, k = 0.49 and
Mo = 0.0722 kg water/kg dry solids. Using these constants, the relationship
between water activity and moisture content was obtained with Eq. (10). The
correlation coefficient between the experimental moisture contents (Table 3)
and the predicted moisture contents from Eq. (10) was 0.993.
2

aw =

0.0722

0.0722
- 1 8.54 - 4(1 - 8.54)
2+
- 1 8.54 - 2 +

Xs
Xs

2( 0.49)(1 - 8.54)

0.5

(10)

Utilizing the saturation vapor pressure of water at 40C obtained from


steam table data as p* = 6980.5 Pa, the RH of the storage environment as

STORAGE, SHELF LIFE AND COLOR OF MANGO POWDER

57

75

Sticky point temperature,Ts, (C)

70

65

60

55

50

0.089

45

40
0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

Moisture content of mango powder, Xp, (kg water/kg dry solids)


FIG. 3. RELATIONSHIP BETWEEN MOISTURE CONTENT AND STICKY-POINT
TEMPERATURE OF MANGO POWDER

TABLE 3.
EXPERIMENTAL MOISTURE CONTENT AND WATER ACTIVITY FOR MANGO POWDER
Time of storage (days)

X (kg water/kg dry solid)

aw

0
15
30
45
60
75
90
105

0.041
0.048
0.056
0.062
0.063
0.068
0.078
0.089

0.226
0.272
0.326
0.395
0.429
0.489
0.576
0.668

Rh = 0.9, the surface area of the pouch as Ap = 2 0.155 0.655 = 0.0512 m2,
the water vapor permeability of the pouch as K = 5.4 10-8 kg/m2/day/Pa and
the amount of dry solids present in 20 g of powder as Ws = 0.02(10.004) =
0.0192 kg and Eq. (10), numerical solution of Eq. (4) resulted in the graphical

Moisture content of mango powder (kg water/kg dry solid)

58

S. JAYA and H. DAS

0.1

0.09

0.08

Predicted curve
0.07

Experimental data points

0.06

0.05

0.04
0

20

40

60

80
100
120
Storage period (day)

140

160

180

200

FIG. 4. RELATIONSHIP BETWEEN THE TIME OF STORAGE AND INCREASE IN MOISTURE


CONTENT OF MANGO POWDER

relationship (Fig. 4) between the time of storage and the moisture content of
mango powder (kg water/kg dry mango solids). Figure 4 also presents the
variation of experimental moisture content with time of storage. From Fig. 4,
the predicted time required for the moisture content of the powder to increase
from an initial value of Xi = 0.004 kg water/kg dry solids to its critical value
of Xc = 0.089 kg water/kg dry solids is 114.68 days, whereas the experimental
time required for this increase of moisture content in mango powder was
105 days.
The predicted moisture content obtained from the model and the experimental moisture contents are very similar (Fig. 4). The correlation coefficient
obtained between the experimental and the predicted moisture contents (from
Eq. 4) is 0.981. The mean relative percentage deviation modulus (E%) (Eq. 7)
between the actual and the predicted moisture contents was calculated as
9.22%. Because the mean relative percentage deviation modulus is less than
10%, we conclude that the Eq. (6) is adequate for predicting the moisture
content and the shelf life of mango powder.

STORAGE, SHELF LIFE AND COLOR OF MANGO POWDER

59

Significant caking was observed by Arya et al. (1986) at 3.25% (db)


moisture content in freeze dried watermelon juice powder in aluminum foillaminated polyethylene pouches of 40-micron thickness during storage at 37C
temperature and 22% RH after 3 months. Arya et al. (1986) also observed that
at 5% moisture content and greater, the watermelon juice powder underwent
extensive caking and discoloration. We observed that the stickiness of mango
powder was obtained at 8.9% (db) moisture content. The water activity of the
mango powder at 8.9% (db) moisture content was 0.67. By considering stickypoint moisture content as the critical moisture content of fruit juice powders,
the quality storage life of fruit juice powders can be determined. Additionally,
from the relationship between sticky-point temperature and sticky-point moisture content of the fruit juice powders, the maximum storage temperature can
be predicted.
Color Change of Mango Powder During Storage
The typical yellow color of the mango powder faded gradually during
storage at accelerated storage conditions. Figure 5 presents the increase in
17.07
18

Overall color difference

16

Experimental data

14

Fitted curve

12
10
8
6
4
2
0
0

20

40

60

80

100

120

Storage time (day)


FIG. 5. RELATIONSHIP BETWEEN OVERALL COLOR DIFFERENCE AND STORAGE TIME
OF MANGO POWDER

60

S. JAYA and H. DAS

overall color differences (DE) with increases in storage time. The value of
DE* = 17.07 (Fig. 5). Using this value of DE* and the experimental DE
values, linear plots between (1) (DE* - DE) and storage time; (2) ln(DE* DE) and storage time and (3) 1/(DE* - DE) and storage time were drawn to
identify the plot that fit the equations best and the correlation coefficients. The
maximum correlation coefficient of 0.99 for the plot between ln(DE* - DE)
and storage time was obtained, and the computed value of the rate constant
(k1) was 0.038/day. The correlation coefficient for zero-order and second-order
equations were 0.80 and 0.79, and the corresponding rate constants were 0.15/
day and 0.02/day, respectively. Therefore, we conclude that the change in
color of the mango powder during storage will follow first-order reaction
kinetics.
Kesseler and Fink (1986) obtained similar results for color changes in
sterilized milk stored at atmospheric conditions. Also, the color changes of
concentrated tomato paste during thermal treatments followed first-order reaction kinetics (Barreiro et al. 1997). In this research, the overall color change
of dried mango powder during accelerated storage conditions followed firstorder reaction rate kinetics.

REFERENCES
ARYA, S.S., PREMAVALLI, K.S., SIDDIAH, C.H. and SHARMA, T.R.
1986. Storage behavior of freeze-dried watermelon. J. Food Sci. Technol.
20, 351357.
BARREIRO, J.A., MILANO, M. and SANDOVAL, A. 1997. Kinetics of color
change of double concentrated tomato paste during thermal treatment. J.
Food Eng. 33, 359365.
BHANDARI, B.R., DATTA, N. and HOWES, T. 1997. Problem associated
with spray drying of sugar rich foods. Dry. Technol. 15(2), 671684.
BHANDARI, B.R., SENOUSSI, A., DOMOULIN, E.D. and LEBERT, A.
1993. Spray drying of concentrated fruit juices. Dry. Technol. 11(5), 33
41.
BUHLER, W. and LIEDY, W. 1989. Characterization of product qualities and
its application in drying process development. Chem. Eng. Process.
26(1), 2734.
CFTRI (CENTRAL FOOD TECHNOLOGICAL RESEARCH INSTITUTE).
1990. Mango in India Monograph for Industry, 3rd Ed., Central food
Technological Research Institute, Mysore, India.
COPLEY, M.J., KAUFMAN, V.F. and RASMUSSEN, C.L. 1956. Recent
development in fruit and vegetable powder technology. Food Technol.
13, 589594.

STORAGE, SHELF LIFE AND COLOR OF MANGO POWDER

61

DOWNTON, G.E., FLORES-LUNA, J.L. and KING, C.J. 1982. Mechanism


of stickiness in hygroscopic amorphous powders. Ind. Chem. Fund. 21,
447451.
HAUGAARD, I.S., KRAG, J., PISECKY, J. and WESTERGAARD, V.
1978. Analytical Methods for Dry Milk Powders. Niro Atomizer,
Denmark.
JAMES, R. 1971. Vacuum puff freeze-drying of tropical fruit juices. J. Food
Sci. 36, 906910.
JAYA, S., SUDHAGAR, M. and DAS, H. 2002. Stickiness of food powders
and related physico-chemical properties of food components. J. Food Sci.
Technol. 39(1), 17.
KESSELER, H.G. and FINK, R. 1986. Changes in heated and stored milk
with an interpretation by reaction kinetics. J. Food Sci. 51, 1105.
LABUZA, T.P. 1984. Moisture Sorption: Practical Aspects of Isotherm Measurement and Use. American Association of Cereal Chemists, Minnesota.
LAU, M.H., TANG, J. and SWANSON, B.G. 2000. Kinetics of textural and
color changes in green asparagus during thermal treatment. J. Food Eng.
45, 231236.
LAZAR, M.E., BROWN, A.H., SMITH, G.S., WANG, F.F. and LINDQUIST,
F.E. 1956. Experimental production of tomato powder by spray drying.
Food Technol. 13(3), 129134.
LEMON, C.J., BAKSHI, A.S. and LABUZA, T.P. 1985. Evaluation of moisture sorption isotherm equation part I: Fruits, vegetable, and meat products. LWT 18, 111115.
MASTERS, K. 1985. Spray Drying Handbook, 4th Ed., Longmont Scientific
and Technical, London, England.
MORGAN, A.I. Jr, GINNETTE, L.F., RANDALL, J.M. and GRAHAM, R.P.
1959. Technique for improving instants. Food Eng. 31(9), 8687.
PAP, L. 1995. Production of pure vegetable juice powders of full biological
value. Fruit Proc. 3, 5560.
PAPADAKIS, S.E. and BABU, R.E. 1992. The sticky issues of drying. Dry.
Technol. 10(40), 817837.
PELEG, M. and HOLLENBECH, A.M. 1984. Flow conditioners and anti
caking agents. Food Technol. 41, 91100.
PISECKY, J. 1985. Standards, specifications and test methods for dry milk
products. In Concentration and Drying of Foods (M.C. Diarmuid, ed.)
pp. 203220, Elsevier, New York.
POTTER, N.N. 1978. Food Science. AVI Publishing Company, Inc., Westport,
Connecticut.
RANGANNA, S. 1987. Handbook of Analysis and Quality Control for Fruit
and Vegetable Products. Tata McGraw-Hill Publications, New Delhi,
India.

62

S. JAYA and H. DAS

ROCKLAND, L.B. and BEUCHAT, L.R. 1987. Water Activity: Theory and
Applications to Food, pp. 203204, Marcel Dekker, New York.
ROOS, Y.H. 1995. Glass transition-related physicochemical changes in foods.
Food Technol. 10, 97102.
ROOS, Y.H. and KAREL, M. 1991. Phase transition of amorphous sucrose
and sucrose solution. J. Food Sci. 56, 3849.
ROOS, Y.H., KAREL, M. and KOKINI, J.L. 1996. Glass transition in low
moisture and frozen foods. J. Food Technol. 11, 95107.
SADASIVAM, S. and MANICKAM, A. 1996. Carbohydrates. In Biochemical
Methods pp. 121, New Age International, New Delhi.
SLADE, L. and LEVINE, H. 1991. Beyond water activity: Recent advances
based on an alternative approach to the assessment of food quality and
safety. Crit. Rev. Food Sci. Nutr. 30, 115360.
SLADE, L. and LEVINE, H. 1994. Water and the glass transition dependence
of the glass transition on composition and chemical structures: Special
implications for flour functionality in cookie baking. J. Food Eng. 22,
143188.
WALLACK, D.A. and KING, C.J. 1988. Sticking and agglomeration of
hygroscopic, amorphous carbohydrate and food powders. Biotechnol.
Prog. 4(1), 3135.

Você também pode gostar