Вы находитесь на странице: 1из 10

This article appeared in a journal published by Elsevier.

The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elseviers archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

Author's personal copy

Journal of Food Engineering 89 (2008) 232240

Contents lists available at ScienceDirect

Journal of Food Engineering


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

A generalized conjugate model for forced convection drying based on an evaporative kinetics
Maria Valeria De Bonis, Gianpaolo Ruocco *
DITEC, Universit degli studi della Basilicata, Campus Macchia Romana, Potenza 85100, Italy

a r t i c l e

i n f o

a b s t r a c t
A model describing the heat and mass transfer involved in food drying is presented. The aim is to determine the effect of air temperature on the performance of the drying process applied to fresh-cut vegetable slices, but other effects can be easily incorporated in the model. The model allows to disregard one of the most limiting parameters in such modeling, i.e. the average heat and mass transfer coefcients at the food/drying substrate interface, which are generally taken from the literature. Such assumptions are limiting in the sense that they are referred to average transfer conditions and general geometries. The presented model relies upon a nite-element solution of time-dependent differential equations for simultaneous and conjugate heat and moisture transfer in a two-dimensional domain, without any inference in such empiricism. A special formulation for drying kinetic of the substrate is also exploited, and a treatment of the dependence of the properties upon the residual moisture content is included. After proper validation with the available experimental measurements, the numerical solution is discussed by presenting each involved eld variables, emphasizing on the conjugate nature of the drying process. Due to its exibility and generality, the model can be used in common industrial driers optimization, even in the assumption of a laminar ow eld. 2008 Elsevier Ltd. All rights reserved.

Article history: Received 6 August 2007 Received in revised form 28 February 2008 Accepted 4 May 2008 Available online 15 May 2008 Keywords: Forced convection Food dehydration Conjugate model Temperature and moisture evolution Local heat and mass transfer

1. Introduction Drying, or dehydration, is one of the most common methods of preserving food, and involves a complex combination of transport phenomena such as the application of heat and the removal of moisture from a food substrate (Barbosa-Canovas and Vega-Mercado, 1996; Fellows, 2000). Drying systems optimization is still sought nowadays and therefore full understanding of these phenomena can help to improve process parameters and hence product quality, emphasizing on the external and internal process parameters that inuence drying behavior. The former include temperature, velocity and relative humidity of the drying medium (air), while the latter include density, permeability, sorption desorption characteristics and physical substrate properties. Starting from the seminal works by Luikov and Whitaker a vast number of contributions has been reported in the last decades on porous and multi-phase media drying by air convection (Chen and Pei, 1989; Barbosa-Canovas and Vega-Mercado, 1996). But in the past few years the multi-dimensional (distributed, transient) analysis has gained importance, specially for lumped moist products, as a considerable computing power became more available, therefore many such studies could be conducted and nalized.
* Corresponding author. Tel.: +393293606237. E-mail address: gianpaolo.ruocco@unibas.it (G. Ruocco). 0260-8774/$ - see front matter 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.jfoodeng.2008.05.008

Shapes and detailed congurations were explored through a variety of approach, though always appealing to empirical, average (i.e. independent on surface locations) relationships for interface transfer calculations. These limitations affected many of the available works on drying modeling in the last decade, as briey recalled in the following. Wang and Chen (1999) presented a thorough diffusive model of heat and mass transfer in moist media, yet limited to a one-dimensional geometry. Chen et al. (1999) developed a nite element model for coupled heat and mass transfer, to implement the thermal processing of chicken patties in a small convection oven with cooking condition and empirical, nonlinear thermal properties. Dincer and his co-workers have presented a great deal of works on the subject in the past and recently (Kaya et al., 2006), where the simultaneous heat and mass transfer have been studied for the spatial variations of the heat and mass transfer coefcients along the treated surface. Pasta drying was studied by Migliori et al. (2005) on an axisymmetric geometry, followed by De Temmerman et al. (2008) who added a radiation driving force. A niteelement approach was employed by Aversa et al. (2007) in order to optimize the drying process, by accounting for local temperature variation of both air and food physical properties. As drying is eminently a conjugate phenomenon (which means, the transfer of mass and heat is solved simultaneously in both solid and uid phases, and are strongly coupled through evaporation

Author's personal copy

M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232240

233

Nomenclature

a
c cp D Dhvap Ea k K K0 K1 H L; L0 ; L00 M

l x
p _ q R

temperature factor in Eq. (8) (dimensionless) concentration (mol/m3) constant pressure specic heat (J/kg K) diffusivity (m2/s) latent heat of evaporation (kJ/kg) activation energy (kJ/mol) thermal conductivity (W/mK) rate of production of water vapor mass (1/s) reference constant (1/s) in Eq. (8) temperature factor in Eq. (8) (dimensionless) height (m) lengths (m) molecular weight (g/mol) dynamic viscosity (Pa s) air absolute humidity (kg water vapor/kg dry air) pressure (Pa) cooling rate due to evaporation (W/m3) universal gas constant (kJ/mol K)

q
t T u; v u U x,y X

air density (kg/m3) time (s) temperature (K) horizontal and vertical component of air velocity (m/s) air velocity vector (m/s) moisture content, wet basis (kg liquid water/kg substrate) horizontal and vertical coordinate (m) moisture content, dry basis (kg liquid water/kg dry substrate)

Subscripts 0 initial a air l liquid water s substrate, bulk v water vapor

and properties variation on moisture and temperature), an innovative, further approach is to solve a model in which the mass and energy interface uxes vary seamlessly in space and time as the solution of the eld variables. With this objective Oliveira and Haghighi (1997) obtained the temperature and the moisture contours for the drying of wood, but their work was affected by the limitation given by considering a laminar boundary layer ow over the substrate. This approach was later complemented by Murugesan et al. (2001) for a timber block using a full Navier-Stokes formulation for the ow eld, allowing for the buoyancy term. While overcoming the limitations of the boundary layer assumptions, their work was focussed on Nusselt and Sherwood numbers on the exposed substrate surface. For the rst time a full conjugate model of a drying food was presented by De Bonis and Ruocco (2007), yet for a specic exchange conguration (a thin baking product to be dried by an impinging turbulent jet draft), where the focus was on residual local water activity. A similar approach is carried out in the present work for a drying vegetable substrate, by employing a nite-element approach. Residual water and temperature elds are computed locally within the substrate, when this interacts with a forced, laminar air ow. The later assumption allows to focus on the basic aspects of ow transport, focussing upon the vapor and liquid water production/ depletion and transport, which is dealt with by an ad-hoc rst-order irreversible kinetics. Such kinetics is included to solve for transient, two-dimensional ow, temperature and moisture elds. Realistic transfer exchanges are inherently considered that vary with process time and surface location, eliminating the need for empirical heat and mass transfer (averaged) coefcients evaluation. 2. Problem formulation Convection and moisture removal by a bulk, hot air draft is assumed to a model substrate, in this case carrot slices, as reported in Fig. 1. During processing, heat is transferred mainly by convection from air to the products exposed surface, and by conduction from the surface toward the substrate interior. Meanwhile, moisture diffuses outward to the surface, where is vaporized. But if the substrate is water saturated, liquid can be converted into vapor even within the substrate, depending on the heat perturbation front. The water transport mechanisms generally include the motion of liquid water (1) by diffusion caused by differences in the concentration of solutes at the surface and in the interior (Fickian diffu-

sion) and (2) by capillary forces, while the motion of water vapor is (3) by diffusion in air spaces within the substrate caused by vapor pressure gradients (Fellows, 2000). 2.1. Assumptions The following assumptions are considered in this work: (1) the ow is laminar; the dryer is two-dimensional, and a small portion in the vicinity of the product is studied only, for sake of simplicity; (2) due to the adopted ow regime, no body force is accounted for; (3) the thermophysical food properties are moisture-dependent, as reported by Ruiz-Lpez et al. (2004), and reported in Table 1, while the air and water properties are temperaturedependent and are taken from Perry et al. (1997): for sake of simplicity their dependency are not reported in the formulation; (4) the effect of capillary forces is included in liquid water diffusivity; (5) the diffusivity of vapor in the substrate is the same than the diffusivity of liquid water, as implied for example by Braud et al. (2001). The following simplifying assumptions are adopted: (1) the viscous heat dissipation in the drying medium and the heat generation within the moist substrate are neglected;

Fig. 1. Geometry and nomenclature.

Author's personal copy

234

M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232240

Table 1 Food properties functions, dependent on the moisture content X (Ruiz-Lpez et al., 2004) Property Function
3

qs (kg/m )
cps (J/kgK) ks (W/mK) Dls , Dvs (m2/s)

440:001 90X X 1750 23451 X 0:49 0:443 exp0:206X 10 2:8527 10 exp0:2283369X

(2) due to the nature of the interacting species, no diffusion uxes are accounted for in the energy equation; (3) neither shrinkage nor deformation of drying substrate are accounted for. 2.2. Governing equations With reference to the previous assumptions, the governing conservation equations in vector form are enforced to yield for concentration of vapor and liquid water, pressure, velocity and temperature (Bird et al., 2002) in two distinct air and substrate sub-domains:  In the substrate: Continuity, liquid water

various parameters that describe both the inherent water phase conversion and interface conditions. In this paper a modied exponential model of evaporation has been adopted, based on an Arrhenius rst-order irreversible kinetics formulation. Several works have been presenting such an approach, as Panagiotou et al. (1999), Azzouz et al. (2002) and Roberts and Tong (2003). It is seen here that the inherent (volumetric) evaporation physics (Roberts and Tong, 2003) must be joined to interface conditions (Panagiotou et al., 1999; Azzouz et al., 2002), such that the thermal, uid dynamic and concentration regimes could be all be represented in the mass source term. The present work is focussed upon the additional dependence on process temperature variation, so that the basic Arrhenius-type relationship can be modied as follows:

K K 0 eEa =RT K a 1
where

ocl r Dls rcl Kcl ot


Continuity, water vapor

ocv r Dvs rcv Kcv ot


Energy

 K 0 is a reference constant, to be found empirically by matching a parametric numerical analysis with the available experimental/ numerical data for each conguration, meaning that for a given conguration (air velocity/humidity, and geometry) K 0 is held constant in the present model;  the activation energy Ea is taken as 48.7 kJ/mol (Roberts and Tong, 2003);  T is the local substrate temperature;  K1 is the ratio of the process temperature to the reference temperature;  a is a dimensionless temperature factor varying with each different process temperature T a . It is emphasized here that present approach, that couples the _ source terms, heat and mass transfer through the use of the K and q simplies the analysis with respect to the classical Luikovs approach, employed by Oliveira and Haghighi (1997) and Murugesan et al. (2001). 2.4. Initial conditions  For the substrate:

oT _ qs cps r ks rT q ot
 In the drying air: Continuity, water vapor

ocv r Dva rcv u rcv ot


Momentum

qa

  ou u ru rp lr2 u ot

Energy

initially in thermal equilibrium (T T 0 ) with the quiescent ambient air the moisture content is such that

qa cpa

oT r ka rT qa cpa u rT ot

cl0 1000

U 0 qs Ml

2.3. Evaporation cooling and vapor production rates _ can be computed as The cooling rate due to evaporation q

It is also cv0 0;  for the drying air: with reference to Fig. 1, no-slip (u  0) is enforced for the drying air at every solid surface; air ows, with a fully-developed (parabolic) horizontal component ua , through the left inlet at given process temperature T a and absolute humidity xa (as usual, related to the relative humidity) such that

_ Dhvap M l Kcl q

The concept of vapor rate of production Kc is adopted in this paper: a negative source term Kcl (K being the rate of production of water vapor mass per unit volume) is included in Eq. (1), to account for the depletion of liquid water; symmetrically, a positive source term Kcv is included in Eq. (2), to account for the production of water vapor. The motivation of the kinetic-like approach for evaporation will be now briey addressed. The rst notion used to describe a drying process incorporating a single constant K for the combined effect of the various existing transport phenomena was suggested by W.K. Lewis in 1921, as recently recalled by Babalis et al. (2006) while reviewing the leading kinetic formulations. Since then, many variations of the basic equations have been reported in the open literature, based on purely empirical models, directly relating moisture ratio with drying time, and incorporating

cv0 1000

xa q a T a xa 1M l

10

It is also cl0 0. 2.5. Boundary conditions Full continuity is assumed for vapor mass and temperature through the substrates surface, to solve for concentrations and temperature seamlessly across the interface. With reference to Fig. 1, the mass, momentum and thermal boundary conditions (where applicable) are as follows:

Author's personal copy

M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232240

235

 Process inlet x 0; 0 < y < Ha

c v c v0 ;

u ua ;

v 0;

T Ta
L0s and L0s L00 s

11
< x < Lp ;

 Bottom plate, air interface 0 < x < y 0

ocv;l 0; oy

u v 0;

oT 0 oy

12

for concentration, velocity, temperature and pressure gradients resolution in the boundary layer and within the substrates exposed surface, induced by the heating and evaporation. Execution time for t = 18000 s elapsed time has been approximately 20 min on a Pentium Xeon PC (WindowsXP Pro OS, 3.0 GHz, 2 GB RAM). Specic underrelaxation factors have been employed to solve the Navier-Stokes equations in the start-up phase of drying. 3. Results and discussion

 Bottom plate, substrate interface L0s < x < L0s L00 s ; y 0

ocl 0; oy

oT 0 oy

13

3.1. Model validation The available literature data are rather limited in order to validate the model and its numerical treatment, as geometry and ow regimes were always left unspecied and transfer coefcients were assumed from empirical correlations, except in Murugesan et al. (2001) (who dealt with a non-food substrate). However, the experimental average residual moisture reported by Ruiz-Lpez et al. (2004) has been rst compared with the present numerical solution and reported in Fig. 3. A 4 h baking process of a thin carrot slice with Ls 0:06 m and Hs 0:0050 or 0.0075 m (for Data Set 1 and 2, respectively) was congured with the following driving parameters: T a 343 or 323 K (for Data Set 1 and 2, respectively), T0 = 298 K, U 0 0:87, and inlet air relative humidity of 45%. Care was exercised to adapt the present model so that the inlet air velocity ua was 2.0 m/s, but still in the laminar regime, with La 0:20 m and Ha 0:10 m. The reference constant K 0 for the given conguration was 7 103 , while the temperature factor a was found to be 0 and -10 for Data Set 1 and 2, respectively. For Data Set 1 there is a good agreement at the beginning of treatment, while a maximum difference of approximately 15% is detected after 2 h. At the end of drying the measured and computed moisture are again very similar. In Data Set 2 the drying condition are milder therefore the kinetic parameters (in absence of a thickness adjustment) underestimate the measurements, the maximum difference being less than 10% after 3 h. A second such benchmark has been found in the numerical data from Aversa et al. (2007) and reported in Fig. 4. A similar process (baking of a carrot substrate) was congured, with Ls 0:06 m and Hs 0:015, and with the following driving parameters: T a 353, 343 or 333 K (for Data Set 3 to 5, respectively), T0 = 303 K, U 0 0:64, and inlet air relative humidity of 75%. Care was exercised, as well, to adapt the present model so that the inlet

 Upper open surface 0 < x < La ; y Hp

ocv 0; oy

ou ov 0; oy oy

oT 0 oy

14

 Process outlet x La ; 0 < y < Ha

ocv 0; ox

ou 0; ox

v 0;

oT 0 ox

15

Finally, continuity is ensured by enforcing the following positions:  Across the horizontal sub-domains interface L0s < x < L0s L00 s; y Hs

cva cvs ;

ocl 0; oy

u v 0;

Ta Ts

16

 Across the upwind x L0s ; 0 < y < Hs and downwind x L00 s ; 0 < y < Hs (vertical) sub-domains interfaces

cva cvs ;

ocl 0; ox

u v 0;

Ta Ts

17

2.6. Numerical method and additional considerations A nite-element commercial solver has been employed to integrate the partial differential equations system (COMSOL Multiphysics Users Guide, 2007). A preliminary grid independency test has been carried out with 3 different grids of approximately 2000, 4000 and 6000 triangular elements, respectively, and the second grid was selected as the local heat ux across the interface vary less than 2% in all locations with respect to the one computed with the third grid. The mesh was distorted locally (Fig. 2) to allow

Fig. 2. Close-up of distorted grid at the substrate (dimensions in m).

Author's personal copy

236

M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232240

Fig. 3. Average U evolution during process: comparison with Ruiz-Lpez et al. (2004) measurements for 2 different Data Sets.

Fig. 4. Average U evolution during process: comparison with Aversa et al. (2007) computations for 3 different Data Sets.

air velocity ua was 0.3 m/s, in the laminar regime, with La 0:20 m and Ha 0:10 m. The reference constant K 0 for the given conguration was 90, while the temperature factor a was found to be 0, 17 and 10 for Data Set 3 to 5, respectively. The same limitations with the earlier benchmark were found, as no information was available on employed conguration. In all

cases a good agreement is detected between the two different models. Small discrepancies (less than 5%) are found after 1 hr of treatment only, due to the condensation phenomenon reported in the benchmark work, which remains unjustied for empirical transfer coefcients such as the ones reportedly employed in Aversa et al. (2007).

Author's personal copy

M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232240

237

3.2. Flow and temperature eld The simulation results for Data Sets 3 conguration (Aversa et al., 2007) are then briey presented in the form of velocity, temperature or moisture distributions. Fig. 5 shows rst the vector and scalar distributions of velocity in the drying air. Due to the ow eld contraction and speed-up, the action of the drying air is strongest on top of the substrate, while the front and back faces are subject to stagnation and recirculation ow regions, respectively. This justies the adoption of a fully conjugate model for a detailed description, as transfer properties vary considerably with exposed surface location. Depending on the ow eld, the temperature distribution in Fig. 6 presents a related non-homogeneous behavior, due to the nonuniform heat transfer, which will then reect upon the residual moisture distribution. On the three exposed substrate sides, due to the conjugate nature of the model, the isotherms are obviously inclined. The substrate is found to be more than 3 K warmer on the leading edge, with respect to the trailing edge, and its left side is being heated more effectively (as expected) than the right one. The lowest temperature of about 340 K is detected on substrate bottom, by the adiabatic oor, with the slowest heating point being located slightly in the ow direction.

In addition to the available Data Set 3, the present model has been exercised by varying the nominal value of velocity. Fig. 7 shows the new ow eld generated with ua = 0.3 m/s, 10 times higher. The velocity distribution is very similar to the previous one, but the velocity local values are much higher indeed. These in turn reect on the higher thermal regime, reported in Fig. 8, where the product center temperature increase by 4 K with respect to Data Set 3 comparison. A more dynamic ow situation dictates an overall more even side-to-side treatment, therefore the slowest heating point is almost perfectly centered this time. 3.3. Moisture and vapor removal Based on the above ow eld and temperature maps, it is expected that (1) the evaporation occurs non-homogeneously within the substrate, and (2) the vapor mass transfer across the uid-substrate interface also occurs non-uniformly. Consequently, the moisture will be non-homogeneously removed within the substrate. Residual moisture distribution after the treatment is reported in Fig. 9 for Data Set 3. The evaporation and depletion of water is more effective where the temperature is the highest (Fig. 6) at the leading edge (a triangular chunk, one-fth of the entire prod-

Fig. 5. Close-up of ow eld (vector eld and streamlines) in the vicinity of substrate for Data Sets 3 to 5 uid dynamic conguration, after a 5 h drying. juj values range from 0 to 0.21 m/s.

Fig. 6. Close-up of temperature eld (isotherms) in the substrate and its vicinity for Data Sets 3 conguration, after a 5 h drying. T values range from 341 to 352 K.

Author's personal copy

238

M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232240

Fig. 7. Close-up of ow eld (vector eld and streamlines) in the vicinity of substrate for a higher air velocity, after a 5 h drying. juj values range from 0 to 2.7 m/s.

Fig. 8. Close-up of temperature eld (isotherms) in the substrate and its vicinity for a higher air velocity, after a 5 h drying. T values range from 345 to 352 K.

Fig. 9. Close-up of residual moisture concentration eld (isolines) in the substrate for Data Set 3 conguration, after a 5 h drying. cl values range approximately from 3.19 to 3.33 104 mol/m3.

uct), but the trailing edge is dried more than the average too, due to the favorable momentum transport in its vicinity.

Fig. 10 describes the effect of ten-fold velocity increment on residual moisture. The drying process is stronger, so the humidity

Author's personal copy

M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232240

239

Fig. 10. Close-up of residual moisture concentration eld (isolines) in the substrate for a higher air velocity, after a 5 h drying. cl values range approximately from 3.10 to 3.23 104 mol/m3.

Fig. 11. Close-up of vapor excess eld (contours) in the vicinity of substrate for Data Set 3 conguration, after a 5 h drying. Representation in 10 levels of gray, for a cv range from approximately 6.05 to 6.20 mol/m 3.

distribution decrease accordingly when compared with the Data Sets 3. The side-to-side treatment is slightly more homogeneous, but a larger concentration is detected y-wise in turn. Finally, Fig. 11 shows the removal of water vapor from the substrate in the uid phase. The much higher vapor concentration in the substrate is not reported in the Figure, for sake of clarity. The vapor is non-uniformly blown away by the air ow. It is interesting to note that such irregular eld, together with the associated moisture, that was presented by one of the benchmark adopted here (Aversa et al., 2007), cannot be predicted under the assumption of average heat and mass transfer coefcients as claimed in their work. On a general basis, an uneven distribution can be computed by adopting local transfer values, as in Kaya et al. (2006), but again this work does not consist in a conjugate model, as the transfer coefcient are empirically implied.

Acknowledgement This work was funded by MIUR Italian Ministry of Scientic Research, Grant No. 2006093719002 entitled Transport phenomena of heat and mass from plates and modied surfaces by air impinging jets. References
Aversa, M., Curcio, S., Calabr, V., Iorio, G., 2007. An analysis of the transport phenomena occurring during food drying process. Journal of Food Engineering 78, 922932. Azzouz, S., Guizani, A., Jomaa, W., Belgith, A., 2002. Moisture diffusivity and drying minetic equation of convective drying of grapes. Journal of Food Engineering 55, 323330. Babalis, S.J., Papanicolau, E., Kyriakis, N., Belessiotis, V.G., 2006. Evaluation of thinlayer drying models for describing drying kinetics of gs (Ficus carica). Journal of Food Engineering 75, 205214. Barbosa-Canovas, G.V., Vega-Mercado, H., 1996. Dehydration of Foods. Chapman & Hall, New York. Bird, R.B., Stewart, W.E., Lightfoot, E.N., 2002. Transport Phenomena. John Wiley & Sons, New York. Braud, L.M., Moreira, R.G., Castell-Perez, M.E., 2001. Mathematical modeling of impingement drying of corn tortillas. Journal of Food Engineering 50, 121128. Chen, P., Pei, D.C.T., 1989. A mathematical model of drying processes. International Journal of Heat and Mass Transfer 32 (2), 297310. Chen, H., Marks, B.P., Murphy, R.Y., 1999. Modeling coupled heat and mass transfer for convection cooking of chicken patties. Journal of Food Engineering 42, 139 146. COMSOL Multiphysics Users Guide, COMSOL AB. 2007. De Bonis, M.V., Ruocco, G., 2007. Modelling local heat and mass transfer in food slabs due to air jet impingement. Journal of Food Engineering 78, 230237. De Temmerman, J., Verboven, P., Delcour, J.A., Nicola, B., Ramon, H., 2008. Drying model for cylindrical pasta shapes using desorption isotherms. Journal of Food Engineering 86, 414421. Fellows, P.J., 2000. Food Processing Technology. CRC Press, Boca Raton. pp. 311317. Kaya, A., Aydn, O., Dincer, I., 2006. Numerical modelling of heat and mass transfer during forced convection drying of rectangular moist objects. Journal of Food Engineering 49, 30943103.

4. Conclusions In this work a generalized conjugate model of forced convection drying has been proposed. A modied exponential model for drying kinetics has been adopted, based on an Arrhenius rst-order irreversible formulation. The basic Arrhenius-type relationship has been modied with an appropriate temperature factor, to deal with the process temperature variations, and validated against the available experimental or numerical literature data. Such an approach is independent on empirical heat and mass transfer coefcients. The proposed model can be complemented by additional air velocity and humidity variation factors, also taking into account of different multi-physics effects such as microwave or ultrasound exposure, and can be readily extended to allow for full threedimensional geometries.

Author's personal copy

240

M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232240 Perry, R.H., Green, D.W., Maloney, J.O., 1997. Perrys Chemical Engineers Handbook. McGraw-Hill, New York. Roberts, J.S., Tong, C.H., 2003. Drying kinetics of hygroscopic porous materials under isothermal conditions and use of a rst-order reaction kinetic model for predicting drying. International Journal of Food Properties 6, 355367. Ruiz-Lpez, I.I., Crdova, A.V., Rodrguez-Jimenes, G.C., Garca-Alvarado, M.A., 2004. Moisture and temperature evolution during food drying: effect of variable properties. Journal of Food Engineering 63, 117124. Wang, Z.H., Chen, G., 1999. Heat and mass transfer during low intensity convection drying. Chemical Engineering Science 54, 38993908.

Migliori, M., Gabriele, D., de Cindio, B., Pollini, C.M., 2005. Modelling of high quality pasta drying: mathematical model and validation. Journal of Food Engineering 69, 387397. Murugesan, K., Suresh, H.N., Seetharamu, K.N., Aswatha Narayana, P.A., Sundararajan, T., 2001. A theoretical model of brick drying as a conjugate problem. International Journal of Heat and Mass Transfer 44, 40754086. Oliveira, L.S., Haghighi, K., 1997. Finite element modeling of grain drying. In: Turner, I., Mujumdar, A.S. (Eds.), Mathematical Modeling and Numerical Techniques. Marcel Dekker, New York, pp. 309338. Panagiotou, N.M., Stubos, A.K., Bamopoulos, G., Maroulis, Z.B., 1999. Drying kinetics of a multicomponent mixture of organic solvents. Drying Technology 17 (10), 21072122.

Вам также может понравиться