Sie sind auf Seite 1von 15

Computers and Electronics in Agriculture

34 (2002) 129143 www.elsevier.com/locate/compag

Theoretical and experimental investigation of


the thermal destruction of Vitamin C in food
pouches
A.G.A. Ghani *, M.M. Farid, X.D. Chen
Department of Chemical & Materials Engineering, The Uni6ersity of Auckland, Pri6ate Bag 92019,
Auckland, New Zealand

Abstract

In this study, theoretical analysis and experimental measurements for the determination of
Vitamin C (ascorbic acid) concentration during thermal sterilization of a three-dimensional
pouch filled with carrot-orange soup, are presented. Experimental validation was performed
by measuring the concentration of ascorbic acid in the soup, using high performance liquid
chromatography (HPLC) and a titration method. Transient temperature, velocity profiles
and the shape of the slowest heating zone (SHZ) during heating were also studied. The
simulation covered a heating cycle of 3000 s. The computational fluid dynamics (CFD) code
PHOENICS was used for this purpose. Saturated steam at a temperature of 121 C (250 F)
was assumed to be the heating medium, which is the sterilization temperature commonly
used in the canning industry. Partial differential equations describing the conservation of
mass, momentum and energy were solved numerically, together with the concentration
equation for Vitamin C, using a finite volume method. The temperature dependence of the
carrot-orange soups viscosity and density was incorporated in the simulation. The results of
the simulations showed that natural convection plays an important role in the heat transfer
within the liquid food in the pouch. The SHZ was found to migrate towards the bottom of
the pouch into a region within 3040% of the pouch height, closest to its deepest end. These
results also showed that the vitamin profiles depend not only on the temperature distribu-
tion, but on the velocity profiles in the pouch. The measured values were compared with
those predicted, showing reasonable agreement with an average deviation of 4.5%. 2002
Elsevier Science B.V. All rights reserved.

Keywords: CFD; Temperature distribution; Velocity profiles; Vitamin C concentration

* Corresponding author. Fax: + 64-9-373-7463.


E-mail address: a.ghani@auckland.ac.nz (A.G.A. Ghani).

0168-1699/02/$ - see front matter 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 1 6 8 - 1 6 9 9 ( 0 1 ) 0 0 1 8 3 - 1
130 A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143

1. Introduction

Chemical and biochemical reactions in liquid food during heating are tempera-
ture-dependent. Such reactions not only destroy microorganisms but also destroy
some of the valuable nutrients, such as vitamins. The destruction of vitamins
follows a first-order reaction similar to microbial destruction (Fellows, 1996).
Because such destruction processes are all temperature-dependent, the distribution
of temperature results in unavoidable spatial distributions of the reaction products.
These distributions can be easily quantified in the case of conductive heating of
solid materials. However, the situation becomes quite complex when the fluid is in
motion. In this case, temperature and concentration profiles are influenced strongly
by the natural convection in liquid food (Fellows, 1996).
The effect of the heat sterilization process for canned foods on their quality and
nutrient retention has been a major concern in thermal processing of food since the
beginning of the canning industry. This concern led to several studies on computer
simulations of food to investigate the effect of thermal sterilization on the retention
of vitamins. Teixeira et al. (1969, 1975) developed a numerical computer model to
simulate the effect of a heat process on the destruction of thiamin in a pea puree,
which is usually relatively rich in thiamin. Saguy and Karel (1979) investigated and
developed a method for calculating the optimum temperature profile as a function
of the time needed to achieve a specified level of sterilization with maximum
nutrient retention.
The research interest in the thermal sterilization of pouches was stimulated by its
approval for use in commercial sterilization of low acid foods. Numerous articles
were published on the benefit of using pouches in food processing (Mermelstein,
1978; Lampi, 1980). Sterilization of food in cans has been well studied both
experimentally and theoretically since the 1920s. Although the pouch was intro-
duced to the market in the early 1960s, little information is available on the
temperature distribution within it during the sterilization process. Castillo et al.
(1980) developed a model to predict the retention of nutrients with first-order
kinetics of thermal degradation in foods packaged in retortable pouches and
sterilized by thermal conduction. The model was effective in predicting the temper-
ature at the center of the container at the end of the heating period. However, the
developed model can be applied only to conduction heating which is limited to solid
food. Tandon and Bhowmik (1986) have done some simulation for liquid-filled
pouch, but based on two-dimensional geometry, which does not represent the true
situation, since the shape of the pouch requires three-dimensional analysis.
The objective of this work was to use CFD technique to provide a rigorous
analysis of the sterilization of a three-dimensional pouch containing liquid foods,
which has not been done previously. The transient temperature, velocity and
concentration profiles of Vitamin C, and the migration of the SHZ during natural
convection heating of food in a 3-D pouch were studied. The concentrations of
ascorbic acid after heat treatment of pouches filled with the liquid food were
measured, to validate the theoretical analysis. The governing equations describing
the fluid motion in the pouch were solved together with the equations describing the
A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143 131

destruction kinetics of bacteria and vitamins. A commercial fluid dynamics analysis


software package, PHOENICS [Concentration, Heat and Mass Limited (CHAM),
London], was used to solve the governing equations because of its availability on
the UNIX computer machine at the University of Auckland. The PHOENICS code
is based on the finite volume method, as developed by Patankar and Spalding
(1972). The details of this code can be found in the PHOENICS manuals, especially
the PHOENICS Input Language (PIL) manual (PHOENICS, 1991), and the
PHOENICS beginners menu system user guide by Radosavljevic and Wu (1990).

2. Numerical approximations and model parameters

2.1. Computational grid

The computations were performed for a three-dimensional pouch, as used by


Ghani et al. (2001a), with a width of 0.12 m, height of 0.04 m and length of 0.22
m. These represent the most probable dimensions of the pouch when lying
horizontally on the tray of the retort. The pouch outer surface temperature (top,
bottom and sides) was assumed to rise instantaneously and to be maintained at 121
C throughout the heating period. The pouch volume was divided into 6000 cells:
20 in the x-direction, 10 in the y-direction and 30 in the z-direction as shown in
Fig. 1. The total simulation time for sterilization of carrot-orange soup was 3000 s.
This required 7 h of CPU time on the UNIX IBM RS6000 workstations at the
University of Auckland.

Fig. 1. Pouch geometry and grid mesh.


132 A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143

Body fitted coordinates (BFC) were used in order to construct the geometry of
the pouch. For generating a curvilinear grid within the sub-domain command, the
option of solving differential equations for the corner coordinates within the
currently active domain was used. This option involves the solution of Laplace
equations for the Cartesian coordinates of the cell corners. In the simulations, a
variety of grid sizes and time steps were used. The results obtained showed that the
solution reported in this study is almost time-step independent and weakly depen-
dent on grid size.

2.2. Go6erning equations and boundary conditions

The partial differential equations governing natural convection motion in a


pouch space are the Navier Stokes equations in x, y and z coordinates (Bird et al.,
1976). These equations were solved together with that for vitamin concentration
given below.
Mass balance for 6itamin

(Cr (C (C (C
+6x r + 6y r +6z r = kTCr (1)
(t (x (y (z

where, 6x, 6y and 6z are the velocities in x, y and z coordinates, respectively (m s 1),
kT is the reaction rate constant of Vitamin C (s 1), Cr is the relative concentration
(%) of Vitamin C in the pouch at any time and location. The concentration is taken
as a dimensionless species concentration, which is defined as the ratio of real time
concentration (C) (kg m 3) to the initial concentration (Co) (kg m 3) multiplied by
100.
The boundary conditions used were: T = Tw, u =0, 6= 0 and w= 0 at top
surface, bottom surface and side walls. The initial conditions used are: Crv = 0,
T =Tref =20 C, u =0, 6 =0 and w =0, where T is the temperature of the liquid
food (C), Tw is the pouch wall temperature (C) and Tref is the food initial
temperature (C).
It is very difficult to measure local vitamin C concentration in the pouch.
Average values of the concentration can be measured by shaking the pouch at the
end of different sterilization times. Hence, for the purpose of comparison with the
theoretical predictions, it is necessary to calculate the mean Vitamin C concentra-
tion in the whole pouch, as described by Ghani et al. (2001b), for can heating. The
weighted average concentration is defined as follows:

n m l
1
C= % % % Ci, j,k DxDyDz (2)
Vpouch i = 1j = 1k = 1

where C( is the weighted average concentration (kg m 3), Vpouch is the volume of
the pouch (m 3), and Dx, Dy and Dz are the spatial increments in the three
directions, as implemented in the software, using a FORTRAN code.
A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143 133

2.3. Physical properties

In the simulations presented here, the viscosity was assumed to be a function of


temperature, following a polynomial form as required by PHOENICS. The viscos-
ity of carrot-orange soup was measured, using a Paar Physica Viscometer VT2
(Anton Paar Gmbh, Graz, Austria), available at the Chemical Engineering Depart-
ment at the University of Auckland. The values of the viscosity were taken at the
extreme low shear rate, which closely simulated the situation in the pouch being
sterilized. These values were 9.79, 5.14 and 2.82 Pa s at 30, 50 and 70 C,
respectively, which was measured based on the decimal accuracy of 9 1% of
maximum value.
The properties of carrot-orange soup used in the current simulations were
calculated from the values reported for the material used in the soup, using their
known mass fractions. The calculated physical properties were: z= 1026 kg m 3,
CP = 3880 J kg 1 K 1, k =0.596 W m 1 K 1 and i= 0.0002 K 1 (Hayes, 1987;
Rahman, 1995), where z, CP, k and i are the density (kg m 3), specific heat
capacity (J kg 1 K 1), thermal conductivity (W m 1 K 1) and thermal expansion
coefficient (K 1) of the liquid food, respectively.
The variation of the density with temperature is usually expressed (Adrian, 1993)
as:

z= zref[1 i(T Tref)] (3)

where Tref and zref are the temperature and density at the reference condition (20
C). The density was assumed constant in the governing equations except in the
buoyancy term (Boussinesq approximation), where Eq. (3) was used to describe its
variation with temperature.

2.4. Vitamin destruction kinetics

The rate of vitamin destruction is usually assumed to follow first-order kinetics.


It is known that the reaction rate constant is a function of temperature and is
usually described by the Arrhenius equation:

kT = A exp
 
E
(4)
RT
where A is the pre-exponential factor (s 1), E is the activation energy [mol 1], R
is the gas constant [kJ (kg mol) 1 K 1]. The kinetic parameters such as reaction
rate constant and activation energy are required for the prediction of quality
changes during food processing (Ilo and Berghofer, 1999). In food process engineer-
ing, the decimal reduction time (D) is used more often. Its relationship is (Heldman
and Hartel, 1997):
2.303
kT = (5)
D
134 A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143

Using the reported decimal reduction time (D), the reaction rate constant kT is
calculated using Eq. (5). Eq. (4) was then used to calculate the pre-exponential
factor A, giving the value of 2.85 109 s 1.
2.5. Assumptions used in the numerical simulation

To simplify the problem, the following assumptions were made:


1. Heat generation due to viscous dissipation is negligible. This is due to the use
of a high viscous liquid, which leads to only low liquid velocities in the pouch.
2. The assumption of no-slip condition at the inside wall of the pouch is valid;
3. The thermal boundary conditions are applied to liquid boundaries rather than
the outer boundaries of the pouch. This is true since the pouch is composed of
very thin layers of polypropylene, aluminium and polyester with a total
thickness less than 0.1 mm;
4. The diffusion effect of ascorbic acid is negligible due to the dominant effect of
convection, and due to the high viscosity of the liquid food used in the
simulation;
5. The condensing steam maintains a constant temperature at the pouch outer
surface;
6. Although the fluid used is non-Newtonian, it is considered as a Newtonian fluid
due to the low shear rate experienced in this application.

3. Experimental analysis of Vitamin C

Five pouches filled with carrot-orange soups were sterilized with different heating
periods 10, 20, 30, 40 and 50 min, respectively. The sterilization was conducted at
the Research and Development Laboratories at Heinz Watties Australasia. After
each sample had been sterilized for a specific period, duplicate samples were taken
from each of the five pouches in addition to the samples taken from a non-sterilized
pouch. The samples were sealed and kept in a refrigerator with sufficient light
protection. The analysis of the Vitamin C was conducted at the Food Science
Laboratory at the Department of Chemistry, University of Auckland, for a
non-sterilized pouch and the five sterilized pouches.
Two methods were used in the determination of Vitamin C concentration, an
HPLC method and 2,6-dichlorophenolindophenol titrimetric method. The HPLC
method was tried and not used after preliminary tests. This was due to the complex
nature of the soup tested, which led to the recording of up to 12 separate peaks in
some of the soup samples.
Measurement of Vitamin C (ascorbic acid) by direct titration with 2,6-
dichlorophenolindophenol dye was used in this study because it is simple and rapid
(Albrecht and Schafer, 1990). Ascorbic acid from a test sample was extracted by
meta-phosphoric acid acetic acid solution, followed by titration with 2,6-
dichlorophenolindophenol dyestuff until a salmon pink color is obtained. The
extracting solution used was 3% meta-phosphoric acid in 0.01 N H2SO4, and the
preparation of the dyestuff solution was based on an official AOAC (1980) method.
A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143 135

Fig. 2. Calibration curve of Vitamin C (ascorbic acid) for the standard samples of 0.067, 0.050 and 0.013
mg/ml.

The method used to extract ascorbic acid from the soup samples was as follows:
1. Weigh 10 g of soup (90.1 g) into a blender.
2. Add 40 ml of the extracting solution.
3. Homogenize the solution by blending for 23 min.
4. Centrifuge the homogenate for 4 5 min at 3000 rpm.
5. Filter the resulting homogenate in a Buchner funnel under vacuum.
6. Use a further 5 ml of extracting solution to rinse out the blender.
7. Transfer the solution to a 50-ml volumetric flask and make up to the mark
using the extracting solution, then take 10 ml aliquots for titration.
Three standard samples of ascorbic acid were used in the extracting solution, in
order to construct its calibration curve. These were 0.067, 0.050 and 0.013 mg/ml.
For titration, three aliquot portions were taken from each sample, and titrated
with the dyestuff solution until a salmon pink coloration, persisting for at least 5 s,
was observed. The volumes required in the titration of the three standard samples
of Vitamin C were 11, 8.2 and 2 ml, respectively. These values were plotted to
construct the calibration curve, as shown in Fig. 2.
Soup samples taken from the pouches, which were sterilized for periods of 0, 10,
20, 30, 40 and 50 min were used in the titration. The arithmetic mean volumes of
dyestuff solution required for these samples were 5.9, 5.7, 5.5, 5.2, 4.7 and 4.3 ml,
respectively. From the calibration curve (Fig. 2), these are equivalent to the
concentrations of 0.072, 0.071, 0.068, 0.064, 0.06 and 0.053 mg/ml, respectively
(Fig. 3).

4. Results and discussions

The effect of sterilization temperature on the rate of the destruction of Vitamin


C is presented here. Figs. 4 and 5 show the results of the simulation at times of 200,
1000 and 3000 s for the temperature distribution and Vitamin C concentration. At
136 A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143

the early stage of heating (t = 200 s), heat transfer is solely controlled by conduc-
tion as shown in Fig. 4a. The destruction of Vitamin C shown in Fig. 5a follows
similar trends.
At later stages of heating, the temperature and concentration profiles of Vitamin
C are seen strongly influenced by convection as shown in Fig. 4b,c and Fig. 5b,c.
The results of the simulation after relatively longer periods of 1000 and 3000 s,
show that most Vitamin C destruction occurs at locations near the wall and
especially at the narrowest end of the pouch, the locations of high temperature. Fig.
5b,c also shows that the vitamins are concentrated at the SHZ. The liquid and thus
the vitamins carried with it are exposed to much less thermal treatment at the SHZ
than the rest of the product.
The buoyancy force created by the change in the density of liquid due to
temperature variation (from walls, top and bottom to the core) produces an upward
flow. The hot liquid going up is deflected by the top wall, and then travels towards
the core. Being heavier, the liquid in the core moves downward and then towards
the wall. Thus a circulating flow is created. As heating progresses, a more uniform
temperature is obtained, reducing the buoyancy forces in the liquid that led to a
significant reduction in the velocity. Fig. 5b shows clearly the combined effects of
the temperature distribution (Fig. 4b) and velocity profile (Figs. 68) on the shape
and location of the high vitamin concentration zone (HVCZ). Fig. 8 for the z-plane
velocity vector, at 8 cm from the widest end, shows the effect of fluid circulation
that leads to the formation of two stagnant zones, which affect the location of
HVCZ, as shown in Fig. 4b. These results lead to the conclusion that the vitamin
profiles depend not only on the temperature distribution, but also on the velocity
profiles in the pouch. The locations of the high vitamin concentration zone occur
almost within the stagnant zones, which correspond to minimum liquid velocity.
Fig. 9 shows the relative concentration profiles of Vitamin C in the x-plane of the
center of the pouch at the end of heating. This figure shows the locations of HVCZ
at the two stagnant zones, which can be explained in terms of the x-plane velocity
vector shown in Fig. 4.

Fig. 3. Experimental destruction of Vitamin C (ascorbic acid) concentration with time.


A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143 137

Fig. 4. Temperature profile planes at 30% height from the bottom of pouch filled with carrot-orange
soup and heated for periods of (a) 200 s; (b) 1000 s; and (c) 3000 s.

Fig. 10 shows the locations of HVCZ at different z-planes in the pouch after
1000 s of heating. Fig. 10a,b shows that the HVCZ occurs in the lower half of the
pouch, which is due to the effect of natural convection currents, such as those
shown in Fig. 4. At locations close to the narrowest end, Fig. 10c,d shows that the
138 A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143

HVCZ stays at the mid-height of the plane. This is due to the influence of
conduction dominating heating in the locations at this end of the pouch.
The average relative vitamin concentration in the whole pouch was calculated,
based on the weighted average of the local concentrations as described by Eq. (2).
The main objective of these calculations was to compare these values with the

Fig. 5. Relative concentration profiles of Vitamin C at 30% height from the bottom of pouch filled with
carrot-orange soup and heated by condensing steam after periods of (a) 200 s; (b) 1000 s; and (c) 3000
s.
A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143 139

Fig. 6. The center of x-plane velocity vector of carrot-orange soup in a pouch heated by condensing
steam after 1000 s, showing the effect of natural convection.

Fig. 7. The y-plane velocity vector of carrot-orange soup in a pouch heated by condensing steam after
300 s, showing the effect of natural convection.

measured one. The theoretical predictions show that at the end of heating (t= 3000
s), the average relative Vitamin C concentrations drop to about 76%. The measured
and predicted relative concentrations of Vitamin C are shown in Fig. 11.
140 A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143

5. Conclusions

It was possible to predict the transient temperature, velocity profile and Vitamin
C destruction during the sterilization of liquid food in pouches. The results of the
simulation show the migration of the slowest heating zone towards the bottom of
the pouch as heating proceeds. The concentration of Vitamin C has been seen to
depend on both temperature distribution and flow pattern. The Vitamin C concen-

Fig. 8. The z-plane velocity vector of carrot-orange soup in a pouch heated by condensing steam after
1000 s, showing the effect of natural convection.

Fig. 9. Relative concentration profiles of Vitamin C at 50% of x-plane of pouch filled with carrot-orange
soup and heated by condensing steam after 3000 s.
A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143 141

Fig. 10. Relative concentration profiles of Vitamin C at different y-planes in a pouch filled with
carrot-orange soup and heated by condensing steam after 1000 s.

tration profiles computed show that the locations of the high vitamin concentration
zone (HVCZ) occur almost within the stagnant zones, where the minimum liquid
velocities prevail. The measured values show reasonable agreement with those
predicted.
142 A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143

Fig. 11. Experimental and theoretical destruction of relative Vitamin C concentration (%) with time.

Acknowledgements

The authors thank Dr Paul Kilmartin, Food Analyst in Food Science Depart-
ment, University of Auckland, for his valuable guidance and advice in the Vitamin
C measurements.

References

Adrian, B., 1993. Heat Transfer. Wiley, New York.


Albrecht, J.A., Schafer, H.W., 1990. Comparison of two methods of ascorbic acid determination in
vegetables. Journal of Liquid Chromatography 13 (13), 2633 2641.
AOAC, 1980. Official Methods of Analysis, 13th Edition. Association of Official Analytical Chemists,
Washington, DC.
Bird, R.B., Stewart, W.E., Lightfoot, E.N., 1976. Transport Phenomena. Wiley, New York.
Castillo, P.F., Barreiro, J.A., Salas, G.R., 1980. Prediction of nutrient retention in thermally processed
heat conduction food packaged in retortable pouches. Journal of Food Science 45, 1513.
Fellows, P.J., 1996. Food Processing Technology (Principles and Practice). Woodhead Publishing Series
in Food Science and Technology, UK.
Ghani, A.G., Farid, M.M., Chen, X.D., Richards, P., 2001a. Thermal sterilization of canned food in a
3-D pouch using computational fluid dynamics. Journal of Food Engineering 48, 147 156.
Ghani, A.G., Farid, M.M., Chen, X.D., Richards, P., 2001b. A CFD study on the effect of sterilization
temperature on bacteria deactivation and vitamin destruction. Journal of Process Engineering 215
(Part E), 9 17.
Hayes, 1987. Food Engineering Data Handbook, Wiley, New York.
Heldman, D.R., Hartel, R.W., 1997. Principles of Food Processing. Chapman & Hall, New York.
Ilo, S., Berghofer, E., 1999. Kinetics of color changes during extrusion cooking of maize gritz. Journal
of Food Engineering 39, 73 80.
Lampi, R.A., 1980. Retort pouch: The development of a basic packaging concept in todays high
technology era. Food Processing Engineering 4, 1 18.
Mermelstein, N.H., 1978. Retort pouch earns 1978 IFT food technology industrial achievement award.
Food Technology 32 (6), 22.
Patankar, S.V., Spalding, D.B., 1972. A calculation procedure for heat, mass and momentum transfer in
three-dimensional parabolic flows. International Journal of Heat and Mass Transfer 15 (10),
1787 1806.
A.G.A. Ghani et al. / Computers and Electronics in Agriculture 34 (2002) 129143 143

PHOENICS Reference Manual, 1991, Part A: PIL, Concentration Heat and Momentum Limited,
CHAM/TR 200 A, Bakery House, London.
Radosavljevic, D., Wu, J.Z., 1990. The PHOENICS beginners menu system user guide. CHAM/TR 217
A, Bakery House, London.
Rahman, R., 1995. Food Properties Handbook. CRC Press, Boca Raton, FL.
Saguy, I., Karel, M., 1979. Optimal retort temperature profile in optimizing thiamin retention in
conduction-type heating of canned foods. Journal of Food Science 44, 1485 1490.
Tandon, S., Bhowmik, S.R., 1986. Evaluation of thermal processing of retortable pouches filled with
conduction-heated foods considering their actual shapes. Journal of Food Science 51 (3), 709 714.
Teixeira, A.A., Dixon, J.R., Zahradnik, J.W., Zinsmeister, G.E., 1969. Computer optimization of
nutrient retention in thermal processing of conduction-heated foods. Food Technology 23 (6),
134 140.
Teixeira, A.A., Stumbo, C.R., Zahradnik, J.W., 1975. Experimental evaluation of mathematical and
computer models for thermal process evaluation. Journal of Food Science 40, 653 655.

Das könnte Ihnen auch gefallen