You are on page 1of 9

Journal of Food Engineering 89 (2008) 232–240

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, Università 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

Article history: A model describing the heat and mass transfer involved in food drying is presented. The aim is to deter-
Received 6 August 2007 mine the effect of air temperature on the performance of the drying process applied to fresh-cut vegeta-
Received in revised form 28 February 2008 ble slices, but other effects can be easily incorporated in the model. The model allows to disregard one of
Accepted 4 May 2008
the most limiting parameters in such modeling, i.e. the average heat and mass transfer coefficients at the
Available online 15 May 2008
food/drying substrate interface, which are generally taken from the literature. Such assumptions are lim-
iting in the sense that they are referred to average transfer conditions and general geometries. The pre-
Keywords:
sented model relies upon a finite-element solution of time-dependent differential equations for
Forced convection
Food dehydration
simultaneous and conjugate heat and moisture transfer in a two-dimensional domain, without any infer-
Conjugate model ence in such empiricism.
Temperature and moisture evolution A special formulation for drying kinetic of the substrate is also exploited, and a treatment of the
Local heat and mass transfer 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 field variables, emphasizing on the conjugate nature of the drying process. Due to its flexibility
and generality, the model can be used in common industrial driers’ optimization, even in the assumption
of a laminar flow field.
Ó 2008 Elsevier Ltd. All rights reserved.

1. Introduction Shapes and detailed configurations were explored through a vari-


ety of approach, though always appealing to empirical, average
Drying, or dehydration, is one of the most common methods of (i.e. independent on surface locations) relationships for interface
preserving food, and involves a complex combination of transport transfer calculations.
phenomena such as the application of heat and the removal of These limitations affected many of the available works on dry-
moisture from a food substrate (Barbosa-Canovas and Vega-Merca- ing modeling in the last decade, as briefly recalled in the following.
do, 1996; Fellows, 2000). Drying systems optimization is still Wang and Chen (1999) presented a thorough diffusive model of
sought nowadays and therefore full understanding of these phe- heat and mass transfer in moist media, yet limited to a one-dimen-
nomena can help to improve process parameters and hence prod- sional geometry. Chen et al. (1999) developed a finite element
uct quality, emphasizing on the external and internal process model for coupled heat and mass transfer, to implement the ther-
parameters that influence drying behavior. The former include mal processing of chicken patties in a small convection oven with
temperature, velocity and relative humidity of the drying medium cooking condition and empirical, nonlinear thermal properties.
(air), while the latter include density, permeability, sorption– Dincer and his co-workers have presented a great deal of works
desorption characteristics and physical substrate properties. on the subject in the past and recently (Kaya et al., 2006), where
Starting from the seminal works by Luikov and Whitaker a vast the simultaneous heat and mass transfer have been studied for
number of contributions has been reported in the last decades on the spatial variations of the heat and mass transfer coefficients
porous and multi-phase media drying by air convection (Chen along the treated surface. Pasta drying was studied by Migliori
and Pei, 1989; Barbosa-Canovas and Vega-Mercado, 1996). But in et al. (2005) on an axisymmetric geometry, followed by De Temm-
the past few years the multi-dimensional (distributed, transient) erman et al. (2008) who added a radiation driving force. A finite-
analysis has gained importance, specially for lumped moist prod- element approach was employed by Aversa et al. (2007) in order
ucts, as a considerable computing power became more available, to optimize the drying process, by accounting for local temperature
therefore many such studies could be conducted and finalized. variation of both air and food physical properties.
As drying is eminently a conjugate phenomenon (which means,
* Corresponding author. Tel.: +393293606237. the transfer of mass and heat is solved simultaneously in both solid
E-mail address: gianpaolo.ruocco@unibas.it (G. Ruocco). and fluid phases, and are strongly coupled through evaporation

0260-8774/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jfoodeng.2008.05.008
M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232–240 233

Nomenclature

a temperature factor in Eq. (8) (dimensionless) q air density (kg/m3)


c concentration (mol/m3) t time (s)
cp constant pressure specific heat (J/kg K) T temperature (K)
D diffusivity (m2/s) u; v horizontal and vertical component of air velocity (m/s)
Dhvap latent heat of evaporation (kJ/kg) u air velocity vector (m/s)
Ea activation energy (kJ/mol) U moisture content, wet basis (kg liquid water/kg sub-
k thermal conductivity (W/mK) strate)
K rate of production of water vapor mass (1/s) x,y horizontal and vertical coordinate (m)
K0 reference constant (1/s) in Eq. (8) X moisture content, dry basis (kg liquid water/kg dry sub-
K1 temperature factor in Eq. (8) (dimensionless) strate)
H height (m)
L; L0 ; L00 lengths (m) Subscripts
M molecular weight (g/mol) 0 initial
l dynamic viscosity (Pa s) a air
x air absolute humidity (kg water vapor/kg dry air) l liquid water
p pressure (Pa) s substrate, bulk
q_ cooling rate due to evaporation (W/m3) v water vapor
R universal gas constant (kJ/mol K)

and properties variation on moisture and temperature), an innova- sion) and (2) by capillary forces, while the motion of water vapor
tive, further approach is to solve a model in which the mass and is (3) by diffusion in air spaces within the substrate caused by va-
energy interface fluxes vary seamlessly in space and time as the por pressure gradients (Fellows, 2000).
solution of the field variables. With this objective Oliveira and Hag-
highi (1997) obtained the temperature and the moisture contours 2.1. Assumptions
for the drying of wood, but their work was affected by the limita-
tion given by considering a laminar boundary layer flow over the The following assumptions are considered in this work:
substrate. This approach was later complemented by Murugesan
et al. (2001) for a timber block using a full Navier-Stokes formula- (1) the flow is laminar; the dryer is two-dimensional, and a
tion for the flow field, allowing for the buoyancy term. While over- small portion in the vicinity of the product is studied only,
coming the limitations of the boundary layer assumptions, their for sake of simplicity;
work was focussed on Nusselt and Sherwood numbers on the ex- (2) due to the adopted flow regime, no body force is accounted
posed substrate surface. For the first time a full conjugate model for;
of a drying food was presented by De Bonis and Ruocco (2007), (3) the thermophysical food properties are moisture-dependent,
yet for a specific exchange configuration (a thin baking product as reported by Ruiz-López et al. (2004), and reported in Table
to be dried by an impinging turbulent jet draft), where the focus 1, while the air and water properties are temperature-
was on residual local water activity. dependent and are taken from Perry et al. (1997): for sake
A similar approach is carried out in the present work for a drying of simplicity their dependency are not reported in the
vegetable substrate, by employing a finite-element approach. formulation;
Residual water and temperature fields are computed locally within (4) the effect of capillary forces is included in liquid water
the substrate, when this interacts with a forced, laminar air flow. diffusivity;
The later assumption allows to focus on the basic aspects of flow (5) the diffusivity of vapor in the substrate is the same than the
transport, focussing upon the vapor and liquid water production/ diffusivity of liquid water, as implied for example by Braud
depletion and transport, which is dealt with by an ad-hoc first-or- et al. (2001).
der irreversible kinetics. Such kinetics is included to solve for tran-
sient, two-dimensional flow, temperature and moisture fields. The following simplifying assumptions are adopted:
Realistic transfer exchanges are inherently considered that vary
with process time and surface location, eliminating the need for (1) the viscous heat dissipation in the drying medium and the
empirical heat and mass transfer (averaged) coefficients evaluation. heat generation within the moist substrate are neglected;

2. Problem formulation

Convection and moisture removal by a bulk, hot air draft is as-


sumed 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 product’s exposed surface, and by conduction from
the surface toward the substrate interior. Meanwhile, moisture dif-
fuses outward to the surface, where is vaporized. But if the sub-
strate 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 concen-
tration of solutes at the surface and in the interior (Fickian diffu- Fig. 1. Geometry and nomenclature.
234 M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232–240

Table 1 various parameters that describe both the inherent water phase
Food properties functions, dependent on the moisture content X (Ruiz-López et al., conversion and interface conditions.
2004)
In this paper a modified exponential model of evaporation has
Property Function been adopted, based on an Arrhenius first-order irreversible kinet-
qs (kg/m3) 440:001 þ 90X ics formulation. Several works have been presenting such an ap-
X
cps (J/kgK) 1750 þ 2345ð1þX Þ proach, as Panagiotou et al. (1999), Azzouz et al. (2002) and
ks (W/mK) 0:49 þ 0:443 expð0:206XÞ Roberts and Tong (2003). It is seen here that the inherent (volu-
10
Dls , Dvs (m2/s) 2:8527  10 expð0:2283369XÞ
metric) evaporation physics (Roberts and Tong, 2003) must be
joined to interface conditions (Panagiotou et al., 1999; Azzouz et
al., 2002), such that the thermal, fluid dynamic and concentration
(2) due to the nature of the interacting species, no diffusion
regimes could be all be represented in the mass source term.
fluxes are accounted for in the energy equation;
The present work is focussed upon the additional dependence
(3) neither shrinkage nor deformation of drying substrate are
on process temperature variation, so that the basic Arrhenius-type
accounted for.
relationship can be modified as follows:

2.2. Governing equations K ¼ K 0 eEa =RT K a1 ð8Þ


where
With reference to the previous assumptions, the governing con-
servation equations in vector form are enforced to yield for concen-  K 0 is a reference constant, to be found empirically by matching a
tration of vapor and liquid water, pressure, velocity and parametric numerical analysis with the available experimental/
temperature (Bird et al., 2002) in two distinct air and substrate numerical data for each configuration, meaning that for a given
sub-domains: configuration (air velocity/humidity, and geometry) K 0 is held
constant in the present model;
 In the substrate: Continuity, liquid water  the activation energy Ea is taken as 48.7 kJ/mol (Roberts and
ocl Tong, 2003);
þ r  ðDls rcl Þ ¼ Kcl ð1Þ  T is the local substrate temperature;
ot
 K1 is the ratio of the process temperature to the reference
Continuity, water vapor
temperature;
ocv  a is a dimensionless temperature factor varying with each dif-
þ r  ðDvs rcv Þ ¼ Kcv ð2Þ
ot ferent process temperature T a .
Energy
It is emphasized here that present approach, that couples the
oT heat and mass transfer through the use of the K and q_ source terms,
qs cps þ r  ðks rT Þ ¼ q_ ð3Þ
ot simplifies the analysis with respect to the classical Luikov’s ap-
 In the drying air: Continuity, water vapor proach, employed by Oliveira and Haghighi (1997) and Murugesan
et al. (2001).
ocv
þ r  ðDva rcv Þ ¼ u  rcv ð4Þ
ot 2.4. Initial conditions
Momentum
   For the substrate:
ou
qa þ u  ru ¼ rp þ lr2 u ð5Þ
ot initially in thermal equilibrium (T ¼ T 0 ) with the quiescent ambi-
ent air the moisture content is such that
Energy
U 0 qs
oT cl0 ¼ 1000 ð9Þ
qa cpa þ r  ðka rT Þ ¼ qa cpa u  rT ð6Þ Ml
ot
It is also cv0 ¼ 0;
2.3. Evaporation cooling and vapor production rates  for the drying air:

with reference to Fig. 1, no-slip (u  0) is enforced for the drying


The cooling rate due to evaporation q_ can be computed as air at every solid surface; air flows, with a fully-developed (par-
q_ ¼ Dhvap M l Kcl ð7Þ abolic) horizontal component ua , through the left inlet at given
process temperature T a and absolute humidity xa (as usual, re-
The concept of vapor rate of production Kc is adopted in this pa- lated to the relative humidity) such that
per: a negative source term Kcl (K being the rate of production of
water vapor mass per unit volume) is included in Eq. (1), to ac-
xa qa ðT a Þ
cv0 ¼ 1000 ð10Þ
count for the depletion of liquid water; symmetrically, a positive ðxa þ 1ÞM l
source term Kcv is included in Eq. (2), to account for the production It is also cl0 ¼ 0.
of water vapor. The motivation of the kinetic-like approach for
evaporation will be now briefly addressed. The first notion used
to describe a drying process incorporating a single constant K for 2.5. Boundary conditions
the combined effect of the various existing transport phenomena
was suggested by W.K. Lewis in 1921, as recently recalled by Full continuity is assumed for vapor mass and temperature
Babalis et al. (2006) while reviewing the leading kinetic formula- through the substrate’s surface, to solve for concentrations and
tions. Since then, many variations of the basic equations have been temperature seamlessly across the interface. With reference to
reported in the open literature, based on purely empirical models, Fig. 1, the mass, momentum and thermal boundary conditions
directly relating moisture ratio with drying time, and incorporating (where applicable) are as follows:
M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232–240 235

 Process inlet ðx ¼ 0; 0 < y < Ha Þ for concentration, velocity, temperature and pressure gradients
resolution in the boundary layer and within the substrate’s ex-
cv ¼ cv0 ; u ¼ ua ; v ¼ 0; T ¼ Ta ð11Þ
posed surface, induced by the heating and evaporation. Execution
 Bottom plate, air interface ð0 < x < L0s and L0s þ L00s < x < Lp ; time for t = 18000 s elapsed time has been approximately 20 min
y ¼ 0Þ on a Pentium Xeon PC (WindowsXP Pro OS, 3.0 GHz, 2 GB RAM).
ocv;l oT Specific underrelaxation factors have been employed to solve the
¼ 0; u ¼ v ¼ 0; ¼0 ð12Þ Navier-Stokes equations in the start-up phase of drying.
oy oy
 Bottom plate, substrate interface ðL0s < x < L0s þ L00s ; y ¼ 0Þ
3. Results and discussion
ocl oT
¼ 0; ¼0 ð13Þ
oy oy 3.1. Model validation
 Upper open surface ð0 < x < La ; y ¼ Hp Þ
The available literature data are rather limited in order to vali-
ocv ou ov oT date the model and its numerical treatment, as geometry and flow
¼ 0; ¼ ¼ 0; ¼0 ð14Þ
oy oy oy oy regimes were always left unspecified and transfer coefficients were
 Process outlet ðx ¼ La ; 0 < y < Ha Þ assumed from empirical correlations, except in Murugesan et al.
(2001) (who dealt with a non-food substrate). However, the exper-
ocv ou oT imental average residual moisture reported by Ruiz-López et al.
¼ 0; ¼ 0; v ¼ 0; ¼0 ð15Þ
ox ox ox (2004) has been first compared with the present numerical solu-
Finally, continuity is ensured by enforcing the following positions: tion and reported in Fig. 3. A 4 h baking process of a thin carrot
slice with Ls ¼ 0:06 m and Hs ¼ 0:0050 or 0.0075 m (for Data Set
 Across the horizontal sub-domains interface ðL0s < x < L0s þ L00s ; 1 and 2, respectively) was configured with the following driving
y ¼ Hs Þ parameters: T a ¼ 343 or 323 K (for Data Set 1 and 2, respectively),
T0 = 298 K, U 0 ¼ 0:87, and inlet air relative humidity of 45%. Care
ocl
cva ¼ cvs ; ¼ 0; u ¼ v ¼ 0; Ta ¼ Ts ð16Þ was exercised to adapt the present model so that the inlet air
oy velocity ua was 2.0 m/s, but still in the laminar regime, with
 Across the upwind ðx ¼ L0s ; 0 < y < Hs Þ and downwind ðx ¼ La ¼ 0:20 m and Ha ¼ 0:10 m. The reference constant K 0 for the gi-
L00s ; 0 < y < Hs Þ (vertical) sub-domains interfaces ven configuration was 7  103 , while the temperature factor a was
found to be 0 and -10 for Data Set 1 and 2, respectively.
ocl
cva ¼ cvs ; ¼ 0; u ¼ v ¼ 0; Ta ¼ Ts ð17Þ For Data Set 1 there is a good agreement at the beginning of
ox treatment, while a maximum difference of approximately 15% is
detected after 2 h. At the end of drying the measured and com-
puted moisture are again very similar. In Data Set 2 the drying con-
2.6. Numerical method and additional considerations dition are milder therefore the kinetic parameters (in absence of a
thickness adjustment) underestimate the measurements, the max-
A finite-element commercial solver has been employed to inte- imum difference being less than 10% after 3 h.
grate the partial differential equations system (COMSOL Multi- A second such benchmark has been found in the numerical data
physics User’s Guide, 2007). A preliminary grid independency from Aversa et al. (2007) and reported in Fig. 4. A similar process
test has been carried out with 3 different grids of approximately (baking of a carrot substrate) was configured, with Ls ¼ 0:06 m
2000, 4000 and 6000 triangular elements, respectively, and the and Hs ¼ 0:015, and with the following driving parameters:
second grid was selected as the local heat flux across the interface T a ¼ 353, 343 or 333 K (for Data Set 3 to 5, respectively),
vary less than 2% in all locations with respect to the one computed T0 = 303 K, U 0 ¼ 0:64, and inlet air relative humidity of 75%. Care
with the third grid. The mesh was distorted locally (Fig. 2) to allow was exercised, as well, to adapt the present model so that the inlet

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


236 M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232–240

Fig. 3. Average U evolution during process: comparison with Ruiz-López 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 ¼ 0:20 m cases a good agreement is detected between the two different
and Ha ¼ 0:10 m. The reference constant K 0 for the given configu- models. Small discrepancies (less than 5%) are found after 1 hr of
ration was 90, while the temperature factor a was found to be 0, treatment only, due to the condensation phenomenon reported
17 and 10 for Data Set 3 to 5, respectively. in the benchmark work, which remains unjustified for empirical
The same limitations with the earlier benchmark were found, as transfer coefficients such as the ones reportedly employed in Aver-
no information was available on employed configuration. In all sa et al. (2007).
M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232–240 237

3.2. Flow and temperature field In addition to the available Data Set 3, the present model has
been exercised by varying the nominal value of velocity. Fig. 7
The simulation results for Data Sets 3 configuration (Aversa shows the new flow field generated with ua = 0.3 m/s, 10 times
et al., 2007) are then briefly presented in the form of velocity, tem- higher. The velocity distribution is very similar to the previous
perature or moisture distributions. Fig. 5 shows first the vector and one, but the velocity local values are much higher indeed. These
scalar distributions of velocity in the drying air. Due to the flow in turn reflect on the higher thermal regime, reported in Fig. 8,
field contraction and speed-up, the action of the drying air is stron- where the product center temperature increase by 4 K with respect
gest on top of the substrate, while the front and back faces are sub- to Data Set 3 comparison. A more dynamic flow situation dictates
ject to stagnation and recirculation flow regions, respectively. This an overall more even side-to-side treatment, therefore the slowest
justifies the adoption of a fully conjugate model for a detailed heating point is almost perfectly centered this time.
description, as transfer properties vary considerably with exposed
surface location. 3.3. Moisture and vapor removal
Depending on the flow field, the temperature distribution in Fig.
6 presents a related non-homogeneous behavior, due to the non- Based on the above flow field and temperature maps, it is ex-
uniform heat transfer, which will then reflect upon the residual pected that (1) the evaporation occurs non-homogeneously within
moisture distribution. On the three exposed substrate sides, due the substrate, and (2) the vapor mass transfer across the fluid-sub-
to the conjugate nature of the model, the isotherms are obviously strate interface also occurs non-uniformly. Consequently, the
inclined. The substrate is found to be more than 3 K warmer on the moisture will be non-homogeneously removed within the
leading edge, with respect to the trailing edge, and its left side is substrate.
being heated more effectively (as expected) than the right one. Residual moisture distribution after the treatment is reported in
The lowest temperature of about 340 K is detected on substrate Fig. 9 for Data Set 3. The evaporation and depletion of water is
bottom, by the adiabatic floor, with the slowest heating point being more effective where the temperature is the highest (Fig. 6) at
located slightly in the flow direction. the leading edge (a triangular chunk, one-fifth of the entire prod-

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

Fig. 6. Close-up of temperature field (isotherms) in the substrate and its vicinity for Data Sets 3 configuration, after a 5 h drying. T values range from 341 to 352 K.
238 M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232–240

Fig. 7. Close-up of flow field (vector field 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 field (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 field (isolines) in the substrate for Data Set 3 configuration, 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 Fig. 10 describes the effect of ten-fold velocity increment on
to the favorable momentum transport in its vicinity. residual moisture. The drying process is stronger, so the humidity
M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232–240 239

Fig. 10. Close-up of residual moisture concentration field (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 field (contours) in the vicinity of substrate for Data Set 3 configuration, 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 Acknowledgement


Sets 3. The side-to-side treatment is slightly more homogeneous,
but a larger concentration is detected y-wise in turn. This work was funded by MIUR Italian Ministry of Scientific Re-
Finally, Fig. 11 shows the removal of water vapor from the sub- search, Grant No. 2006093719002 entitled ‘‘Transport phenomena
strate in the fluid phase. The much higher vapor concentration in of heat and mass from plates and modified surfaces by air imping-
the substrate is not reported in the Figure, for sake of clarity. The ing jets.”
vapor is non-uniformly blown away by the air flow. It is interesting
to note that such irregular field, together with the associated mois- References
ture, that was presented by one of the benchmark adopted here
(Aversa et al., 2007), cannot be predicted under the assumption Aversa, M., Curcio, S., Calabrò, V., Iorio, G., 2007. An analysis of the transport
phenomena occurring during food drying process. Journal of Food Engineering
of average heat and mass transfer coefficients as claimed in their 78, 922–932.
work. On a general basis, an uneven distribution can be computed Azzouz, S., Guizani, A., Jomaa, W., Belgith, A., 2002. Moisture diffusivity and drying
by adopting local transfer values, as in Kaya et al. (2006), but again minetic equation of convective drying of grapes. Journal of Food Engineering 55,
323–330.
this work does not consist in a conjugate model, as the transfer Babalis, S.J., Papanicolau, E., Kyriakis, N., Belessiotis, V.G., 2006. Evaluation of thin-
coefficient are empirically implied. layer drying models for describing drying kinetics of figs (Ficus carica). Journal
of Food Engineering 75, 205–214.
Barbosa-Canovas, G.V., Vega-Mercado, H., 1996. Dehydration of Foods. Chapman &
Hall, New York.
4. Conclusions Bird, R.B., Stewart, W.E., Lightfoot, E.N., 2002. Transport Phenomena. John Wiley &
Sons, New York.
In this work a generalized conjugate model of forced convection Braud, L.M., Moreira, R.G., Castell-Perez, M.E., 2001. Mathematical modeling of
impingement drying of corn tortillas. Journal of Food Engineering 50, 121–128.
drying has been proposed. A modified exponential model for dry-
Chen, P., Pei, D.C.T., 1989. A mathematical model of drying processes. International
ing kinetics has been adopted, based on an Arrhenius first-order Journal of Heat and Mass Transfer 32 (2), 297–310.
irreversible formulation. 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–
The basic Arrhenius-type relationship has been modified with
146.
an appropriate temperature factor, to deal with the process tem- COMSOL Multiphysics User’s Guide, COMSOL AB. 2007.
perature variations, and validated against the available experimen- De Bonis, M.V., Ruocco, G., 2007. Modelling local heat and mass transfer in food
tal or numerical literature data. Such an approach is independent slabs due to air jet impingement. Journal of Food Engineering 78, 230–237.
De Temmerman, J., Verboven, P., Delcour, J.A., Nicolaï, B., Ramon, H., 2008. Drying
on empirical heat and mass transfer coefficients. model for cylindrical pasta shapes using desorption isotherms. Journal of Food
The proposed model can be complemented by additional air Engineering 86, 414–421.
velocity and humidity variation factors, also taking into account Fellows, P.J., 2000. Food Processing Technology. CRC Press, Boca Raton. pp.
311–317.
of different multi-physics effects such as microwave or ultrasound Kaya, A., Aydın, O., Dincer, I., 2006. Numerical modelling of heat and mass transfer
exposure, and can be readily extended to allow for full three- during forced convection drying of rectangular moist objects. Journal of Food
dimensional geometries. Engineering 49, 3094–3103.
240 M.V. De Bonis, G. Ruocco / Journal of Food Engineering 89 (2008) 232–240

Migliori, M., Gabriele, D., de Cindio, B., Pollini, C.M., 2005. Modelling of high quality Perry, R.H., Green, D.W., Maloney, J.O., 1997. Perry’s Chemical Engineers’ Handbook.
pasta drying: mathematical model and validation. Journal of Food Engineering McGraw-Hill, New York.
69, 387–397. Roberts, J.S., Tong, C.H., 2003. Drying kinetics of hygroscopic porous materials under
Murugesan, K., Suresh, H.N., Seetharamu, K.N., Aswatha Narayana, P.A., isothermal conditions and use of a first-order reaction kinetic model for
Sundararajan, T., 2001. A theoretical model of brick drying as a conjugate predicting drying. International Journal of Food Properties 6, 355–367.
problem. International Journal of Heat and Mass Transfer 44, 4075–4086. Ruiz-López, I.I., Córdova, A.V., Rodríguez-Jimenes, G.C., García-Alvarado, M.A., 2004.
Oliveira, L.S., Haghighi, K., 1997. Finite element modeling of grain drying. In: Turner, Moisture and temperature evolution during food drying: effect of variable
I., Mujumdar, A.S. (Eds.), Mathematical Modeling and Numerical Techniques. properties. Journal of Food Engineering 63, 117–124.
Marcel Dekker, New York, pp. 309–338. Wang, Z.H., Chen, G., 1999. Heat and mass transfer during low intensity convection
Panagiotou, N.M., Stubos, A.K., Bamopoulos, G., Maroulis, Z.B., 1999. Drying kinetics drying. Chemical Engineering Science 54, 3899–3908.
of a multicomponent mixture of organic solvents. Drying Technology 17 (10),
2107–2122.

You might also like