Anda di halaman 1dari 8

Journal of Food Engineering 111 (2012) 613

Contents lists available at SciVerse ScienceDirect

Journal of Food Engineering


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

Toroid cans An experimental and computational study for process innovation


Mustafa Karaduman, Rahmi Uyar, Ferruh Erdogdu
Department of Food Engineering, University of Mersin, Ciftlikkoy-Mersin 33343, Turkey

a r t i c l e

i n f o

a b s t r a c t
Use of toroid cans might be an innovative approach to increase heat transfer rates in canning. Therefore, the objectives of this study were to design a toroid can and to determine its possible use. For these objectives, a toroid can was designed, and water lled and pineapple slices placed cans were thermally processed. Tomato paste lled cans were used to demonstrate the improvements in heating rates. In convection heating, recirculation in both side walls was observed computationally to lead to increased heat transfer rate while the cold point shifted from center towards to outer wall in conduction heating due to the effect of heat transfer from inside surface. Enhancement in heat transfer leads to shorter processing times to achieve required sterility with better quality attributes, and toroid cans might be a possible process innovation in canning with this respective. 2012 Elsevier Ltd. All rights reserved.

Article history: Received 28 August 2011 Received in revised form 18 January 2012 Accepted 5 February 2012 Available online 15 February 2012 Keywords: Canning Toroid cans Heat transfer

1. Introduction Thermal processing continues to be the most effective way in food processing to extend the shelf life of food products for a safe consumption (Abdul Ghani et al., 2003), and canning is still a universal and economic method for processing (Weng, 2005). Heat transfer rate in canning is inuenced by container shape, product type and process temperature (Krishnamurty et al., 2001). Starting from Nicholas Apperts process in glass bottles in 1810, introduction of cylindrical tin cans was a signicant improvement (Cowell, 2007). With different sizes, they have continued to rule. Developments since their inception included soldered or welded cans, two- and three-piece cans and tin-free steel and aluminum cans (Lewis, 2006). As indicated by Brody (2002), metal cans, however, are not necessarily cylindrical any more, and various shapes included at rectangles, tall and thin, squat, pot, pot-bellied, wasp-waist, parallelopiped, multi-ringed, twisted, etc. All these modications were to reduce costs and provide convenience in processing (Lewis, 2006). Due to the complex heat transfer within the modied shapes, temperature patterns should be known to achieve required sterility while minimizing cooking (Brody, 2002) for an optimum process. Geometry modication seems to be one possible way to increase the heat transfer rate since the distance that heat needs to transfer from outer boundary to the lowest lethal heat point (Bown, 2004), As reported by Varma and Kannan (2006), geometry modication and its orientation are to enhance the heat transfer. This is effective in liquid or semi-uid food products when the heat transfer mechanism
Corresponding author. Tel.: +90 533 812 0686; fax: +90 324 361 0032.
E-mail addresses: ferruherdogdu@mersin.edu.tr, ferruherdogdu@yahoo.com (F. Erdogdu). 0260-8774/$ - see front matter 2012 Elsevier Ltd. All rights reserved. doi:10.1016/j.jfoodeng.2012.02.009

is convection (Varma and Kannan, 2005; Varma and Kannan, 2006) computationally investigated the heat transfer in from full through truncated cones to full cylinder cans to reduce the required sterilization time. The focus was to enhance the sterilization process through geometry modications without mechanical agitation or rotation. In fact, agitation equipments were reported not to justify low volume processes (Kannan and Sandaka, 2008). In canning, heat transfer mechanism between containers and heating medium (processing medium of pure steam,steam/air mixture, water spray or water) is convection, and temperature difference is generally xed by process design considerations. Hence, increasing convective heat transfer coefcient (h) and/or increasing surface area and decreasing the center-to-surface distance of the containers are possible ways to increase heat transfer rates for reducing the process time. In the high temperature heat transfer medium (pure steam; steam/air mixture; water spray; water) (Bown, 2004), h is assumed to be high in process design. In addition to the possibilities in the heat transfer medium, the heat transfer rate in liquid or semi-uid food products can be signicantly increased by mechanical agitation. However, it is not the case for conduction-heated foods. This leaves the possibility of increasing surface area for increased heat transfer rate. However, volume increase with surface area results in larger quantity of food product to process leading to longer processing times to achieve the required sterility with an expense to higher cook values and quality losses. It is also important that shape modication is also expected to affect the heat transfer rate along the outer surface especially when the external heat transfer medium is air. Abraham and Sparrow (2003) obtained the convective heat transfer rates at the surface of a thermal load as a function of the enclosure-to-load temperature difference and stated the prevailing of steady heat transfer rates at the load surfaces while Sparrow and Abraham

M. Karaduman et al. / Journal of Food Engineering 111 (2012) 613

(2002) determined the heat transfer coefcients and other performance parameters for variously positioned and supported thermal loads in ovens with/without water lled or empty blockages. Based on these, a design to increase the surface area without further increasing the volume might be a possible way to increase heat transfer rates. Increasing heat transfer rate improves overall quality of the food product by minimizing color, avor and nutrient losses (Price and Bhowmik, 1994). Additional benets would be reduced operating and nal costs due to the reduction in the processing time. Toroid is a doughnut-shape generated by revolving a rectangle 360 to generate a hollow cylinder. A toroid can design might be an innovation in canning to increase the heat transfer rate with additional boundary inside the hollow cylinder. Therefore, the objectives of this study were to design a toroid can to increase the heat transfer rate compared to cylindrical cans and to determine the increased heat transfer rates. 2. Materials and methods For the given objectives, the study consisted of experimental and computational parts. In the experimental part, a toroid can was rst designed, and water lled toroid can was processed in hot water in retort under pasteurization conditions (98.19 0.36 C). In the second part of the experiments, pineapple slices were placed in the toroid can and processed in hot water. In the latter part of the experimental part, heat transfer rates in the tomato paste lled toroid and regular cylindrical cans were compared. In the computational part, the conductive and convective heat transfer within the can were investigated by using water and pineapple samples. In water and pineapple cases, timetemperature data obtained during the process was used to validate the computational models. Resulting temperature distribution was used to demonstrate the higher rate of heat transfer and movement in the cold spot location compared to regular cylindrical cans. Flow patterns, temperature distribution and shapes of the slowest heating zones in these cans were also predicted and compared using the computational models to provide insight on natural convective and conductive heat transfer processes in the toroid can and to better understand the heat transfer phenomena for improving the design and operation for an innovative process. 2.1. Experimental methodology In the experimental part, a custom-made toroid can was designed and constructed (Fig. 1a) using a 1.5 mm thick stainless steel (AISI304). Inner and outer diameters of the hollow structure were 18 and 73 mm with a 110 mm length. Including the waterproof lid, top surface thickness was 4 mm. Given dimensions of the toroid can gave only 6% reduction in the volume compared to the volume of a regular #1 can (73 110 mm) while the side surface boundary increased from 73p to (73 + 18) p mm2 (25% increase). Needle type-T thermocouples were located inside these cans using ring gaskets and locking-receptacles (Fig. 1b). Fig. 2 shows the location of the thermocouple in the toroid can. After locating the thermocouple, the toroid can was lled with distilled water, and the lid was tightly closed for processing in hot water in a retort. Toroid cans seemed to be suitable for processing of pine-apple slices due to their hollow structure (Fig. 3). Hence, they were also processed after the pineapple slices were placed in the toroid can to observe the possible way of using these cans in processing. For this purpose, pineapple was purchased from a local market and located in the toroid can to process in hot water in retort under pasteurization conditions. A regular cylindrical can was also constructed (73 110 mm with a 4 mm thick top lid) using the AISI304 stainless steel

Fig. 1. Custom-made toroid can (a) with the inserted needle type-T thermocouple (b).

material to compare the heating rates. After locating a needle type-T thermocouple in the cylindrical can (Fig. 4), both cans were lled with tomato paste, purchased from a local store, for further processing in hot water. Tomato paste was specically chosen to represent a conduction heated food product. In all cases, timetemperature data were recorded using a Keithley 2700 DMM (coupled with Keithley 7700.20CH multiplexer, Keithley Instruments, Cleveland, OH, USA) data acquisition system, and the results were compared with the regular cans to demonstrate the higher heating rates in the toroid cans. Averaged data from 5 experiments with their standard deviations were used to validate the computational models. 2.2. Computational model A 2-dimensional (2D) axi-symmetric geometry set-up used to describe the toroid shape can with the applied mesh structure is shown in Fig. 5. The given dimensions and wall thicknesses were used through the whole study. The geometry setup to represent the geometrical conguration of the toroid can was prepared using Ansys V12 preprocessor mesh generation interface (Ansys Inc., Canonsburg, PA). As shown, high resolution of mesh congurations consisting of quadrilateral structural mesh cells were used along the wall surfaces and the axi-symmetry line. Since accuracy of the numerical model depended strongly on the mesh resolution in the vicinity of the wall surfaces and the axi-symmetry line, non-uniformly spaced grid points were used to obtain a denser mesh along these surfaces with a high resolution of mesh

M. Karaduman et al. / Journal of Food Engineering 111 (2012) 613

Thermocouple 4

31.1

Thermocouple

110

15 1.5

26.9 55 1.5

110

1.5

1.5
1.5

38 cm
9 29m
Fig. 4. Location of thermocouple in the regular cylindrical can (scaled in mm).

Fig. 2. Location of thermocouple in the toroid can (scaled in mm).

Fig. 3. Toroid can with pineapple slices.

conguration. This was due to the expected strong velocity and temperature gradients in this layer (Rabiey et al., 2007). 2D quadrilateral cells starting from the boundaries were then created to build up the 2D computational domain. The number of mesh cells through the wall thickness (1.5 mm) was 3, and the computational domain was discretized into a nal mesh system of control volumes of 20201 quadrilateral mesh cells in the 2D plane (Fig. 5). Mesh size, time-step size and convergence criteria were all considered together while nalizing the given mesh conguration. 3. Governing equations The required governing equations were continuity, energy and momentum equations for liquid medium while only the solution of the energy equation was used for solid mediums: Continuity equation:

1 @ @ r qf v r qf v z 0 r @r @r

Fig. 5. The applied mesh system of control volumes of 20201 quadrilateral mesh cells in the 2-dimensional plane.

M. Karaduman et al. / Journal of Food Engineering 111 (2012) 613

" # kf @T @T @T 1 @ @T @2T vr vz r 2 @t @r @ z qf cpf r @ r @ r @z   @v r @v r @v r @P vr vz @r @t @r @z " l


Momentum equation in the radial direction:

Energy equation:

Table 1 Temperature-variable thermo-physical properties of water applied in the simulations (engel, 2007). Temperature (K) 293.15 313.15 333.15 353.15 373.15 393.15 Density (kg/m3) 998 992.1 983.3 971.8 957.9 943.4 Viscosity (Pa-s) 0.001002 0.000653 0.000467 0.000355 0.000282 0.000232 Thermal conductivity (W/m-K) 0.598 0.631 0.654 0.670 0.679 0.683 Specic heat (J/kg-K) 4182 4179 4185 4197 4217 4244

qf

@ 1 @ @2v r r v r 2 @r r @r @z

# 3

" #   @v z @v z @v z @P 1 @ @v z @2v z l r qf vr vz 2 qf g @z r @r @t @r @z @r @z 4
where T is temperature (K), P is pressure (Pa), t is heating time (s), g is gravitational acceleration (9.81 m/s2), l is dynamic viscosity (Pas), kf, qf and cpf are the thermal conductivity (W/m-K), density (kg/ m3) and specic heat (J/kg-K) of the uid and vr and vz are the velocity components (m/s) in radial and vertical directions, respectively. Angular direction effects were not involved since the axisymmetric approach was applied in the computations. With this approach, computational time was reduced considerably. In addition, laminar ow mode was assumed through the whole heating process. Rayleigh number, calculated based on a vertical cylinder, was in the range of 109. Further information on this issue was given by Erdogdu et al. (2010). In another study (Kiziltas et al. 2010), laminar assumption in the liquid phase was also applied, and comparison of experimental results with the simulation results indicated the validity of this assumption. All thermal and physical properties (thermal conductivity, specic heat, viscosity and density), as explained below, were applied to be functions of temperature. For the case of pineapple lled toroid can, only the energy equation, excluding the velocity components, was required to solve. Thermal conductivity, specic heat and density of the pineapple were assumed to be constant with the given initial and boundary conditions. 3.1. Boundary conditions No-slip boundary conditions for velocity components and uniform wall temperature were imposed at the top, bottom and side surface inside walls of the ow domain with the axi-symmetric wall boundary condition at the axi-symmetry line. A constant uniform initial temperature (21.76 C) distribution in the toroid can was assumed with an innite heat transfer coefcient due to the effect of hot water (98.19 C) and thin walls (1.5 mm) with higher thermal conductivity value. Hence, a constant surface temperature to dene the thermal boundary conditions was applied at the top, bottom and side surface outer walls. Validity of the innite heat transfer coefcient assumption was demonstrated by Erdogdu et al. (2010). As also reported by Erdogdu and Tutar (2011), no other thermal boundary conditions were required since the applied numerical approach determined the heat transfer directly from the solution in the adjacent mesh cells. Same thermal boundary conditions was applied for the case of pineapple lled toroid can with the constant uniform initial temperature (21.76 C) distribution. 3.2. Thermal and physical properties Temperature dependent thermal conductivity, specic heat, viscosity and density for water phase were applied in the simulations (Table 1). Effect of temperature variable properties compared to the constant properties is rather signicant in modeling the natural convection heat transfer. Abdul Ghani et al. (1999) also addressed

Momentum equation in the vertical direction:

the signicance of temperature variable viscosity compared to a constant viscosity value. The density was also assumed to be the function of temperature to account the natural convection effects. Thermal conductivity, specic heat and density of the stainless steel (AISI304) material were 16.2 W/m-K, 502 J/kg-K and 7900 kg/ m3, respectively. For pineapple, thermal and physical properties were obtained from the literature; thermal conductivity: 0.549 W/m-K; specic heat 3490 J/kg-K; density: 1010 kg/m3 (Abdul Ghani and Farid, 2006). 3.3. Solution methodology Ansys/Flotran (Ansys Inc., Canonsburg, PA) discipline was used for the simultaneous determination of velocity components and temperature changes with the Collocated Galerkin approach in the Advection options to obtain a convergent solution for momentum, pressure and temperature variables. Modied inertial relaxation parameter in the momentum equations was set to 1 to prevent the possible spurious oscillations and convergence difculties. The three diagonal matrix algorithm (TDMA) was used to determine the velocity and temperature changes while preconditioned conjugate gradient method (PCGM) was applied for pressure changes. Higher number of iterations were required in the Ansys/Flotran module to improve the accuracy due to the lower viscoisty of water, and time step of Dt = 103 s was used. A similar comment on the effect of lower viscosity value of water was also noted by Abdul-Ghani et al. (1999). For computations of the pineapple lled toroid can, Ansys/ Thermal discipline was used to determine the temperature changes during th conduction heating. For accuracy,time step of Dt = 102 s was used. Calculations were carried out on a HP Z400 Xeon computer with 2.4 GHz CPU speed and 12 GB RAM. 4. Results and discussion The computational model was rst validated with the experimental data where timetemperature data obtained from the
100 80 60
Experimental

Temperature (C)

40 20 0 0 30 60 90 120 150 180

Model

Time (s)
Fig. 6. Comparison of computational model results for toroid can with the experimental data.

10

M. Karaduman et al. / Journal of Food Engineering 111 (2012) 613

Fig. 7. Velocity variation along the can cross-section and recirculation of liquid along the side walls to increase the heat transfer rate (a) 30 s of processing; (b) 90 s of processing.

Fig. 8. Temperature contours evolved in the toroid can cans - (a) 30 s of processing; (b) 90 s of processing.

water lled toroid can were used. Fig. 6 shows the comparison of computational model results with the experimental data where the averaged data from 5 experiments were used. As observed,

computational model results agreed well with the experimental data validating the 2D axi-symmetric assumption for toroid cans.

M. Karaduman et al. / Journal of Food Engineering 111 (2012) 613

11

Ostrach (1988) discussed a great diversity and a broad range of natural convection based buoyancy ows in enclosures. It was reported that the internal natural convection problems (as presented in this study) are considerably more complex than external ones due to the rather large Rayleigh numbers that might be encountered (Ostrach, 1988). During heating of liquid water by natural convection, part of the liquid in the wall vicinities expands resulting in an increase in the local pressure with signicant effects in heat transfer due to the thermal buoyancy in the gravitational force eld (Aktas and Farouk, 2003). In the case of regular cylindrical cans, this phenomena is seen along the side wall where the liquid receives the heat resulting in its expansion and getting a lower density. This leads to development of an upward buoyancy force with a motion due to the density differences carrying the colder liquid upward by viscous drag (Kiziltas et al., 2010). The liquid owing upward is then deected by the top surface and starts moving in the radial direction. This results in the liquid getting cooler and heavier and starting to move downward. As the liquid descends, its temperature decreases upon mixing with colder layers with creation of a new cycle from bottom leading to ow recirculation and increase in the heat transfer rate compared to the conduction heating. In the case of toroid cans, this phenomenon is seen in both right and left side walls resulting in a more increased rate of heat transfer. Fig. 7 shows the velocity prole and the recirculation in both walls compared to the case in the regular cans where the recirculation and development of the upward buoyancy force with a motion are observed only in the side wall. Temperature contours for these vector elds are also given in Fig. 8 to better demonstrate the effect of natural convection buoyancy force on both left and right walls. As in the regular cylindrical cans, formation of secondary ow (formation of eddies) with Benard convection cells were also observed at the bottom of the toroid can due to the effect of heat transfer from the bottom (Fig. 7). As reported by Abdul Ghani and Farid (2006), random nature of the Benard convection cells caused the irregular shape of the isotherms and ow formation near to bottom surface, and due to the effect of recirculating ow along the side walls, the cold point was also found to be pushed towards the bottom similar to the case of a regular cylindrical can. Fig. 9 shows the comparison of experimental timetemperature data with computational model results for processing pineapple slices in the toroid cans. As observed, the model results agreed well with the experimental data. In a similar manner, Abdul Ghani and Farid (2006) also treated the pineapple slices, located in a regular cylindrical can with liquid juice, as impermeable solid for modeling purposes, and axi-symmetric approach was applied to simulate the timetemperature distribution during the process. Temperature
120 100

Temperature (C)

80 60 40 20 0 0 300 600 900 1200 1500 1800 2100


Experimental Model

Time (s)
Fig. 9. Comparison of computational model results with the experimental data for ineapple slices process in the toroid can. Fig. 10. Temperature kernels of the pineapple slices processed in the toroid cans (a) 600 s of processing; (b) 1800 s of processing.

12

M. Karaduman et al. / Journal of Food Engineering 111 (2012) 613

contours for pineapple processing in the toroid cans are shown in Fig. 10 to represent the temperature kernels observed in the conduction heated foods. Due to the effect of heat transfer from the inside wall, the cold point, which appears at the center in regular cylindrical cans, shifted towards the inner wall from center due to the effect of heat transfer from the inside wall. Location of the cold point and its possible movement are critical parameters in determining the required sterility. In conduction-heated canned foods, it lies at the geometrical center of the can, and the distance from outer boundary is the longest. This leads to longer processing times to achieve the required sterility. Since the distance that needs to travel from outer boundary to the cold point is shortened in toroid can design and the cold point location is shifted as demonstrated in Fig. 10. This is expected to lead to the shorter processing times to achieve the required sterility with a better color, avor and lower nutrient losses due to reduced exposure to higher processing temperatures. To compare the heating rate in the toroid cans with the regular cylindrical cans, both cans were lled with tomato paste, representing a conduction heated food product, and processed in hot water. The resulting timetemperature data were then used to determine the slope of the linear portion of the temperature ratio h i T T 1 lnT :R: ln T versus time curves to represent the heating i T 1 rates where Ti is the initial temperature, T1 is the medium temperature and T is the temperature of the tomato paste. Rabiey et al. (2007) also presented the thermal evolution in term of dimensionh i T 1 T less temperature T in a similar way. Fig. 11a and b show the 1 T i change in the temperature ratio and the linear portion of the temperature ratio in the toroid can compared to the regular can. The averaged results from the three experimental data was 0.0007 s1 for the regular cylindrical can while almost four times

higher value (0.0029 s1) was obtained for the toroid can demonstrating the higher heating rate. 5. Conclusions Canning is still a universal and economic method for preserving and processing. In canning, heat penetration rate is inuenced by container shape, product type and process temperature. Modications in container shape were for cost reductions and to provide convenience in processing. In this perspective, application possibilities of toroid cans were determined experimentally and computationally for conduction and convection heated cases. Temperature distribution inside toroid cans including water as convection heating case and pineapple slices representing a suitable product for processing were determined and experimentally validated. The results demonstrated the signicant effect of additional boundary inside the toroid can to increase the heating rate. The improvement in the heating rate of the toroid can compared to the regular cylindrical can was also demonstrated processing the tomato paste. Hence, toroid cans might be introduced a possible innovation for canning industry, and the economical considerations should be determined in further studies. However, this modication will be different from the commercial practice since a various modications will be required for especially lling and sealing. It should also be noted that making such cans will not be practical due to the possible problems during lling and sealing. As stated by Wang et al. (2009) toroidal surfaces are more complex and evaluation of torisity has not been throughly performed compared with simples forms such as planes, spheres, cylinder and cones. Based on the proven usefulness of toroid shapes with possible increased heat transfer rates, on the other hand, various applications in food processing area might be suggested in addition to canning, e.g., expansion of continuous ow turbo cookers to torus/toroid shape, an extension of helical heat exchangers or optimal design of spiral sterilizers. In a similar manner, to improve the mixing, a time and energy consuming problem commonly encountered in chemical and food industries, Dieulot et al. (2005) investigated a torus model containing a sliding well-mixed zone as a way to represent mixing process at unsteady stirring conditions in agitated vessels while Helbling et al. (2011) presented the mathematical modeling of controlled release of drug delivery from torusshaped single-layer devices. Acknowledgement This study was supported by the Undergraduate Students Research Program (2209-A) of the Scientic and Technical Research Council of Turkey (TUBITAK). References

(a)

0.75

T.R.

0.5

Regular can Toroid can

0.25

0 0 500 1000 Time (s) Time (s) 0 2000 4000 6000 8000 1500 2000

(b) 0
-1
-2

Regular can -3 Toroid can


-4

-5
-6
Fig. 11. Comparison of the heating rate in the toroid cans with the regular cylindrical cans (a) temperature ratio (T.R.) versus time (b) natural logarithm of temperature ratio versus time.

Abdul Ghani, A.G., Farid, M.M., Chen, X.D., Richards, P., 1999. Numerical simulation of natural convection heating of canned food by computational uid dynamics. Journal of Food Engineering 41, 5564. Abdul Ghani, A.G., Farid, M.M., Zarrouk, S.J., 2003. The effect of can rotation on sterilization of liquid food using computational uid dynamics. Journal of Food Engineering 57, 916. Abdul Ghani, A.G., Farid, M.M., 2006. Use of computational uid dynamics to analyze the thermal sterilization of solidliquid food mixture in cans. Journal of Food Engineering 7, 5561. Abraham, J.P., Sparrow, E.M., 2003. Three dimensional laminar and turbulent natural convection in a continuously/discretely wall-heated enclosure containing a thermal load. Numerical Heat Transfer-A 44, 105125. Aktas, M.K., Farouk, B., 2003. Numerical simulation of developing natural convection in an enclosure due to rapid heating. International Journal of Heat and Mass Transfer 46, 22532261. Bown, G., 2004. Modeling and optimizing retort temperature control. In: Richardson, P. (Ed.), Improving the thermal processing of foods. CRC Press Taylor & Francis, Boca Raton, FL, pp. 105123. Brody, A.L., 2002. Food canning in the 21st century. Food Technology 56 (3), 7578.

ln[T.R.]

M. Karaduman et al. / Journal of Food Engineering 111 (2012) 613 Cowell, N.D., 2007. More light on the dawn of canning. Food Technology 61 (5), 40 45. engel, Y.A., 2007. Heat and Mass Transfer, third ed. McGraw Hill Companies, Inc. New York, NY, USA. Dieulot, J.-Y., Petit, N., Rouchon, P., Delaplace, G., 2005. A torus model containing a sliding well-mixed zone as a way to represent mixing process at unsteady stirring conditions in agitated vessels. Chemical Engineering Communications 192, 805826. Erdogdu, F., Tutar, M., 2011. Velocity and temperature eld characteristics of water and air during natural convection heating in cans. Journal of Food Science 76, E119E129. Erdogdu, F., Uyar, R., Palazoglu, T.K., 2010. Experimental comparison of natural convection and conduction heat transfer. Journal of Food Process Engineering 33, 85100. Helbling, I.M., Luna, J.A., Cabrera, M.I., 2011. Mathematical modeling of drug delivery from torus-shaped single-layer devices. Journal of Controlled Release 149, 258263. Kannan, A., Sandaka, P.Ch.G., 2008. Heat transfer analysis of canned food in a still retort. Journal of Food Engineering 88, 213228. Kiziltas, S., Erdogdu, F., Palazoglu, T.K., 2010. Simulation of heat transfer for solid liquid food mixtures in cans 3 and model validation under pasteurization conditions. Journal of Food Engineering 97, 449456. Krishnamurthy, H., Ramaswamy, H.S., Sanchez, G., Sablani, S., Pandey, P.K., 2001. Effect of guar gum concentration, rotation speed, particle concentration, retort temperature, diameter of rotation on heat transfer rates during end-over-end processing of canned particulate non-Newtonian uids. In: Welti-Chanes, J. (Ed.), Proceedings of the Eight International Congress on Engineering and Food

13

(ICEF 8), Puebla, Mexico, vol. I. Technomic Publishing Company, Lancaster, PA, pp. 665670. Lewis, M.J., 2006. Thermal processing. In: Brennan, J.G. (Ed.), Food processing handbook. Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, pp. 3369. Ostrach, S., 1988. Natural convection in enclosures. Journal of Heat Transfer 110, 11751190. Price, R.B., Bhowmik, S.R., 1994. Heat transfer in canned foods undergoing agitation. Journal of Food Engineering 23, 621629. Rabiey, L., Flick, D., Duquenoy, A., 2007. 3D simulation of heat transfer and liquid ow during sterilization of large particles in a cylindrical vertical can. Journal of Food Engineering 82, 409417. Sparrow, E.M., Abraham, J.P., 2002. Heat transfer coefcients and other performance parameters for variously positioned and supported thermal loads in ovens with/ without water-lled or empty blockages. International Journal of Heat and Mass Transfer 45, 35973607. Varma, M.N., Kannan, A., 2005. Enhanced food sterilization through inclination of the container walls and geometry modications. International Journal of Heat and Mass Transfer 48, 37533762. Varma, M.N., Kannan, A., 2006. CFD studies on natural convection heating of canned food in conical and cylindrical containers. Journal of Food Engineering 77, 10241036. Wang, Y., Li, L., Ni, J., Huang, S., 2009. Toroidal surfaces using particle swarm optimization. Journal of Manufacturing Science and Engineering, 131, 051015 1 - 0510159. Weng, Z.J., 2005. Thermal processing of canned foods. In:S. Da-Wen (Ed.), Thermal food processing - new technologies and quality issues (pp. 335362). CRC Press Taylor & Francis, Boca Raton, FL.

Anda mungkin juga menyukai