You are on page 1of 15

Heat Transfer—Asian Research, 28 (5), 1999

Mechanism and Mathematical Model of Heat and Mass Transfer


During Convective Drying of Porous Materials

Zhe Zhang
Institute of Engineering Thermophysics, Chinese Academy of Sciences, P.O. Box 2706, Beijing,
P. R. China 100080
Shiming Yang
Department of Power Machinery Engineering, Shanghai Jiao Tong University, Shanghai,
P. R. China 200030
Dengying Liu
Institute of Engineering Thermophysics, Chinese Academy of Sciences, P.O. Box 2706, Beijing
P. R. China 100080

Based on classical analysis and conclusive comments about various kinds of


drying models, a rigorously formulated and comprehensive theoretical model is
established to describe heat and mass transfer during constant rate and falling rate
periods in convective drying of porous materials. The concept of iterative correction
is introduced, and a corresponding numerical method is developed for the moving
boundary problem in numerical simulations of drying processes. The calculation
results for the drying of bricks show that the model presented is more precise than
other models. © 1999 Scripta Technica, Heat Trans Asian Res, 28(5): 337–351, 1999

Key words: Porous materials, convection, drying, heat and mass transfer

1. Introduction

Convective drying of porous materials is one of the oldest and most common unit operations
found in many different processes in the agricultural, ceramic, chemical, food, mineral, textile, and
wood industries. It is also one of the most complex and least-understood operations. Since the phase
change is often concurrent with the drying process of porous media, every process in the interior of
a porous medium becomes very complicated. Moreover, since the mechanism of moisture migration
is dominated by many physical variables, research on convective drying of moist porous media has
at present not produced satisfactory results. Each researcher sticks to his own view, and the models
are so diverse that the calculation results sometimes do not coincide because different models are used
for the same problem. The classical models of heat and mass transfer in drying problems can be divided
into three categories—coupled-field driving models, continuum models, and combined theoretical
models—according to the main development of heat and mass transport theories in porous media in
the past hundred years as well as the understanding of transport mechanisms by the authors of the
models [1]. Coupled-field driving models include single-field driving, two-field driving, and three-
field driving models. Assuming that the gradient of liquid content, or vapor diffusion, or capillary
CCC1099-2871/99/050337-15
© 1999 Scripta Technica
337
flow was the only mechanism of internal moisture transport in pores, Sherwood [2–3], King [4],
Krischer [5], Berger [6], and Eckert, etc. [7–8] proposed a classical liquid diffusion model, a vapor
diffusion model, and a capillary driving model, respectively, to describe heat and mass transfer in
porous materials, all of which are single-field driving models. In the two-field coupling model first
developed by Philip and De Vries [10, 11], a temperature gradient and moisture gradient were regarded
as driving potentials, and the liquid and vapor mass fluxes were defined by Darcy’s law and the Stefan
diffusion law, respectively. Luikov [12] proposed a three-field driving model utilizing three dependent
variables, namely, temperature, saturation, and pressure, based on irreversible thermodynamics.
Continuum models are mainly based on the fundamental laws of conservation of mass, momentum,
and energy. With the number of governing equations and the described physical quantities increasing,
the application range of this type of model widens. Models established by the continuum approach
can be classified into two divisions, Whitaker’s school and Darcy’s school. The models in Whitaker’s
school, such as Whitaker’s model [14], are based on Whitaker’s multiphase transport theorem [13]
and volume-averaging technology, while in the models of Darcy’s school, the momentum equation
is based on Darcy’s law, which is only modified with some unsaturated coefficients [9, 16]. Broadly
speaking, coupled-field driving models and continuum models both describe continuous fields. They
differ in their methodologies for solving problems. By combining Luikov’s model with Whitaker’s
model, Ilic [17] proposed a model that describes the evolution of temperature, pressure, and moisture
distributions in both wet and dry regions in porous media.

In comparison with other models, Whitaker’s model is more rigorously formulated and
comprehensive, because it establishes the macroscopic equations through a volume-averaging tech-
nique at a control point and considers a great deal of the transport mechanism of moisture and energy
in the media. However, its most important drawback is perhaps the determination of the numerous
transport coefficients, some of which are in the form of second-order tensors in the case of
non-homogeneous and anisotropic media (e.g., effective thermal-conductivity tensor, gas-phase total
effective diffusivity tensor). In most problems, it is difficult, even impossible, to determine some of
the transport coefficients in Whitaker’s model. Note too that all of these models regard the vapor and
liquid phases as moisture, which not only makes it difficult to determine mixture transport coefficients
but also leads to difficulties in analyzing the motion of liquid and vapor in pores. In fact, vapor moves
together with non-condensible gas (air) due to the pressure difference (density difference) in moist
porous media. However, the above models only consider the diffusion of vapor. The movement of
vapor together with air as a gas mixture is not considered. In view of these above-mentioned
drawbacks, we regard the air and vapor in pores as a gas mixture phase, and consider for the vapor
both the movement as a gas mixture phase and diffusion in steady air in this paper. From the continuum
approach, based on irreversible thermodynamics theory in coupled-field driving models, using
Whitaker’s local volume average method, a combined theoretical model of describing flow and heat
transfer in porous materials during convective drying is developed in this paper.

Due to the moving boundary problem in numerical simulations of drying processes, the
computation is rather complex. In order to improve the calculation speed and accuracy, the concept
of iteration correction is also introduced.

Nomenclature

c: specific heat capacity [Jkg–1K–1]

338
Db: bound water conductivity [m2s–1]
Dd : thermal dispersion coefficients [m2s–1]
Dva : vapor diffusion coefficient [m2s–1]
Ea : activation energy of movement of bound water [Jmol–1]
g: gravitational constant [ms–2]
H: height [m]
hm: convective mass transfer coefficient [ms–1]
ho: convective heat transfer coefficient [wm–2K–1]
Jb: bound water flux [kgm–2s–1]
k: intrinsic permeability of porous material [m2]
k′: relative permeability
L: length [m]
.
m: evaporation rate [kgm–3s–1]
p: pressure [Nm–2]
pc: capillary pressure, pg − pl, [Nms–2]
R: gas constant [Jmol–1K–1]
S: saturation
Sav : average saturation
T: temperature [°C]
t: time [s]

V: velocity vector [ms–1]
v: vertical velocity [ms–1]
x,y: space coordinate [m]
yi(t): evaporation front location [m]
α: correction factor in Eq. (30)
β: volumetric expansion coefficient [K–1]
γ: specific heat of vaporization [Jkg–1]
ε: volumetric phase content
λm: apparent thermal conductivity [wm–1K–1]
µ: dynamic viscosity [kgm–1s–1]
ρ: density [kgm–3]
σ: surface tension [Nm–1]
n: kinematic viscosity [m2s–1]
ϕ: porosity

Subscripts

0: initial
a: air
b: bound water
c: critical
g: gas
l: liquid
s: solid
sr: surface
v: vapor

339
2. Mechanism of Drying and Its Mathematical Description

Drying materials may be classified into hygroscopic and non-hygroscopic materials. For
non-hygroscopic porous materials, the water inside the pores that contribute to flow is either free
water or unbound water. However, the water inside very fine capillaries cannot easily be replaced by
air. This portion of water is known as the irreducible water content or maximum sorptive water
content. It is also defined as bound water. The structure of many hygroscopic materials, such as food
products and wood, is basically an array of cells [9]. Although the cells are very similar in nature, the
tissue may differ widely in porosity and bulk properties. The intercellular spaces, like the voids in
porous materials, are interconnected and filled with air and a certain amount of free water. The cells
themselves also contain water, which is also called bound water. The cellular membrane behaves like
a perfectly semi-permeable membrane and may act as a capillary path for bound water to migrate.
When water content is lower than the maximum irreducible water content εl,ir, water will exist as
bound water in the pores. However, when the water content exceeds the maximum irreducible water
content, water will exist as free water and form a water ring, which may be continuous or discontinuous
depending on the water content.

2.1 Drying period

2.1.1 Constant rate drying period

If the initial moisture content of porous materials is high enough, the surface is covered with
a continuous layer of free water (water in funicular state) and evaporation takes place mainly at the
surface. The movement of liquid is not only maintained by capillary force, but also driven by many
other forces, such as liquid-solid matrix interfacial drag (Darcy resistance of liquid phase), the inertial
force of the movement of the liquid film, viscous force and gravity, etc. At this point, internal moisture
transfer to the surface and the evaporation at the surface are in equilibrium, and the free water on the
surface will be evaporated steadily and continuously. Therefore, the drying rate is determined by
external conditions only, and a constant drying rate period will be observed. If the temperature gradient
within the material is negligible, the surface temperature is almost constant, and its value is very close
to the wet bulb temperature of the flowing air.

2.1.2 First falling rate period

As drying proceeds, the fraction of wet area at the surface decreases with decreasing surface
water content. When water passes through randomly distributed paths in a medium, there is a
percolation threshold, which usually corresponds to a critical free water content. When the free water
content is greater than the critical, the water phase is continuous. Regardless of the rate of internal
moisture transfer, as long as the free water content at the surface is less than the critical value (this
critical saturation is about 0.3 for most porous materials [11]), the surface will form discontinuous
wet patches. Thus, the mass transfer coefficient decreases with the surface free water content, and the
first falling rate period starts. In the first falling rate period, a new energy balance will be reached at
the surface, accompanied by slowly rising surface temperature. Free water still exists at the surface,
the dry patches still contain bound water, and the vapor pressure at the surface is determined by the
Clapeyron equation. At this point, the convective mass and heat transfer coefficients at the surface

340
are a function of the surface water content. The experimental results of Nissan et al. [18] show that
the relation between the heat and mass transfer coefficient and the surface water content is as follows,

 εl − εl,ir 
ho = ho,0 ηh + (1 − ηh)  (1)
 εl,c − εl,ir 

 εl − εl,ir 
hm = hm,0 ηm + (1 − ηm)  (2)
 εl,c − εl,ir 

where εl,c is the critical water content, εl,ir is the maximum irreducible water content, and coefficients
ηh and ηm are constants, determined experimentally.

2.1.3 Second failing rate period

When the surface water content reaches the maximum sorptive value after further drying, dry
patches appear and there is no free water on the surface. The surface temperature increases rapidly,
signaling the start of the second falling rate period, during which a receding evaporation front often
appears, dividing the system into two regions, a wet region and a dry region (sorption region), as
shown in Fig. 1. Inside the evaporation front, the material is wet; i.e., the voids contain free water,
and the main mechanisms of moisture transfer are capillary flow, viscous force, and inertial force.
Outside the front, no free water exists, all water is in the sorptive or bound water state and the main

Fig. 1. Drying model.

341
mechanisms of moisture transfer are movement of bound water and vapor transfer. Evaporation takes
place at the front as well as in the whole sorption region, while vapor flows through the sorption region
to the surface. As the amount of water evaporating from the surface decreases quickly, the surface
temperature rises much faster than in the first falling rate period.

Movement of bound water is sometimes known as liquid moisture transfer near dryness or
sorption diffusion. It may flow along very fine capillaries or through cellular membranes. It may have
the characteristics of capillary flow, with the driving force being the gas phase or vapor pressure
gradient, i.e.,

kb ∗ ∂ψ (3)
Jb = −ρl Pv ∇εl = −ρlDb∇εl
µ ∂εl

where Jb is the bound water flux and kb the permeability of bound water; P∗v is the saturated vapor
pressure; ψ is the relative humidity, ψ = Pv / P∗v; Db is defined as the bound water conductivity, which
is an Arrhenius-type function of temperature and can be written as

3
 εl,b − εl,eq   Ea 
Db = Db,0   exp − RT  (4)
 εl,ir − εl,eq   

where εl,eq is the equilibrium water content. It is the minimum water content to which a material can
theoretically be dried under the given process conditions; Ea is defined as the transfer activation energy
of the bound water. At room temperature, bound water conductivity depends on the microscopic
structure of the material and is of the order of 10−9m2s−1. Therefore, in the wet region, it can be
neglected, as compared with the liquid water conductivity of free water, which is of the order of
10−7m2s−1.

2.2 Drying model

Based on the above analysis and the conservation laws, we established the following mathe-
matical model to describe heat and mass transfer during constant rate and failing rate periods in
convective drying of porous media. The physical configuration of the model is shown in Fig. 1.

2.2.1 Wet region model

Mass conservation equation:

Liquid phase:

∂(ρlεl) . (5)
+ ∇ ⋅ (ρlVl) = −m
∂t

Gas phase:

342
∂(ρgεg) . (6)
+ ∇ ⋅ (ρgVg) = m
∂t

Air:

∂(ρaεg) (7)
+ ∇ ⋅ (ρaVg − ρvV gv) = 0
∂t

Vapor:

∂(ρgεg) . (8)
+ ∇ ⋅ [ρv(V gv + Vg)] = m
∂t

.
where the subscripts, s, l, g, v, a represent solid, liquid, gas, vapor, and air, respectively; m is the
→ g
evaporation rate, and V v is the vapor diffusion velocity.

Momentum equation:

Gas phase:

. ngεg (9)
∂Vg mVg 1
+ Vg ⋅ ∇Vg + =− ∇pg − εgg + ng∇2Vg − g Vg
∂t ρg ρg kkg

Liquid phase:

. ngεl (10)
∂Vl mVl
= − ∇pl − εl→
1
+ Vl ⋅ ∇Vl − g + nl∇2Vl − g Vl
∂t ρl ρl kkl

Vapor diffusion equation:

P 1 (11)
V gv = −εgDva ⋅ ∇ρv
P − Pv ρv

Energy equation:

∂T
(εsρscs + εlρlcl + εgρgcg) + (ρlclVl + ρacaVa + ρvcvVv) ⋅ ∇T
∂t
.
+ γm = ∇ ⋅ [(λm + ρlclDd)∇T] (12)

In the constant and the first falling rate drying period, when free water still exists on the surface,
a wet region appears in the whole porous material and the boundary conditions are

343
y = H:

ρvnv + ρlnl = hm(ρv − ρv,a) (13)

∂T
λm + γ(ρlnl + ρvnv) = ho(Ta − T) (14)
∂y

y = 0:

ρvnv + ρlnl = 0 (15)

∂T (16)
λm =0
∂y

According to the Levert equation, capillary pressure is given as

pc = pg − pl = √ ϕk (σ0 − βT)J(S)

where J(S) = 0.364{1 − exp[−40(1 − S)]} + 0.221(1 − S) + 0.005 / (S − Sir).

The pressure gradient of the liquid phase in Eq. (10) can be obtained using the capillary
pressure gradient and gas pressure gradient. The relative permeability of gas and liquid kgg and kgl in
Eq. (9) and Eq. (10) can be calculated by the following equations, respectively,

kgg = 1 − 1.11S, when S < Sc (17)

3
 S − Sir  (18)
kgl = , when S > Sir
 1 − Sir 

The thermal dispersion coefficient Dd in energy equation (12) can be found in Ref. 20.

2.2.2 Dry region model

In the second falling rate drying period, a receding evaporation front often appears, dividing
the system into two regions: the wet region and dry region, as shown in Fig. 1. The governing equations
and boundary conditions in this period are:

Dry region:

Mass conservation equation:

Bound water:

344
∂(ρlεb) . (19)
+ ∇ ⋅ (ρlVb) = −mb
∂t

Air:

∂(ρaεg) (20)
+ ∇ ⋅ (ρaVg − ρvV gv) = 0
∂t

Vapor:

∂(ρgεg) . (21)
+ ∇ ⋅ (ρv(V gv + Vg)) = mb
∂t

where subscript b represents bound water.

Momentum equation:

Gas phase:

. ngεg (22)
∂Vg mVg 1
+ Vg ⋅ ∇Vg + =− ∇pg − εgg + ng∇2Vg − g Vg
∂t ρg ρg kkg

Bound water:

Vb = −Db∇εb (23)

Energy equation:

∂T
(εsρscs + εbρbcb + εgρgcg) + (ρbcbVb + ρacaVa + ρvcvVv) ⋅ ∇T
∂t
.
+ γmb = ∇ ⋅ [λm∇T] (24)

Boundary condition at y = H:

pg = patm , ρvnv + ρlnb = hm(Y − Y∞) (25)

∂T (26)
λm + γ(ρlnb + ρvnv) = ho(Ta − T)
∂y

at the evaporation front at y = yi(t):

345
T(x, t) and Pg(x, t) are continuous, and

 ∂T   ∂T  + − (27)
λm ∂y  = λm ∂y  + γρv{vv[yi (t)] − vv[yi (t)]}
 y+i (t)  y−i (t)

∂yi(t) (28)
ρv{vv[y+i (t)] − vv[y−i (t)]} = ρlvl[y−i (t)] − ρlεϕ
∂t

where vv is the vertical velocity of vapor diffusion.

The governing equations for heat and mass transfer in the wet region are the same as Eqs.
(5)–(12).

In the above dry and wet region models, the balance among such forces as liquid phase
macroscopic inertial force, the change of momentum induced by evaporation, capillary force, body
force, viscous force, and the liquid-solid matrix interfacial drag (Darcyian resistance of liquid phase)
has been taken into consideration in the momentum equation for the liquid phase, Eq. (10). The balance
among inertial force, increase of momentum induced by the generation of vapor, gas phase pressure
gradient, body force, viscous force, and the gas-solid matrix drag force is also expressed in the
momentum equation for the gas phase, Eq. (9). The energy balance among convective heat transfer
of the gas and liquid phases, inner heat transfer by phase change, heat conduction through various
phases, and thermal dispersion has been taken into account in the energy equation. Thus it can be seen
that these equations adequately describe the different transport mechanisms of liquid, vapor, and air
in porous materials. Moreover, by describing more transport mechanisms in drying and by connecting
the movement of liquid and gas mixture and the diffusion of vapor organically, the model compre-
hensively describes the fluid flow and heat transport in the convective drying of porous materials.
Under specific drying condition, this model can calculate the field distributions and their variations
of temperature, phase content, pressure, the migrate velocity of liquid, gas, vapor, and the evaporation
rate during drying.

3. Numerical Calculation Method

For the series of equations mentioned above, we obtained solutions by a finite difference
method. The gas momentum equation and energy equation were solved by the SIMPLE algorithm
[21]. At the second falling rate stage, an adaptive mesh technique was used that allowed for the
concentration of nodes near the moving interface.

In order to improve the accuracy of the calculations and accelerate the iteration, we propose
an iterative correction method according to the overall mass balance law. That is, in a certain time
step (e.g., from t to t + ∆t), the overall mass of water evaporated is equal to the overall mass of water
decreased within porous materials, i.e.,

LH L  H
  . .  
 dx∆t (29)
∫∫ ϕ∆S(x, y)ρldxdy = ∫  ∫
 
m(x, y)dy + m sr (x) 
0  0
 
0 0

346
where ∆S(x, y) represents the change of saturation at position (x, y) within one time step.

We judge the calculation in a certain time step to be convergent or not in terms of the above
overall mass balance. If the results do not satisfy the required accuracy, the surface saturation is
corrected using the following equation:

LH L H    (30)
   . .  dx ∆t
∫∫ ∫ ∫
a
0 (x) = S0(x) −
Si+1 ϕ∆S(x, − ( + (x)
i
y)ρldxdy m x, y)dy m sr   
L 00     
  
0 0
 

and then repeat to solve the governing equations to obtain the results for time t + ∆t. This procedure
continues until convergence is reached. The introduction of this iterative correction method not only
guarantees numerical accuracy, but also lets a larger time step ∆t be chosen to shorten the computa-
tional time.

In order to further improve the calculation accuracy, the derivative on the boundaries is
calculated using three-point-value formulation. For the stagger mesh shown in Fig. 2, it is:

 ∂f  8f (I, M1) − 9f (I, M2) + f (I,M3) (31)



∂y  = 3δx
 M1

In the constant rate period and the first falling rate period, the wet region model was solved
for the velocity of gas mixture, liquid and vapor, saturation, evaporation rate, and temperature field.
The vapor density was found from the Clausius–Clapeyron equation. In the second falling rate period,
wet and dry regions coexist, so the conservation equations in the two domains (dry region and wet
region), and the conditions at the interface of these domains must be solved simultaneously. In the
dry (sorptive) region, the unknowns are temperature, gas phase pressure, vapor density, evaporation
rate and migration velocity of bound water, and bound water content.

Fig. 2. Boundary grid.

347
When the surface water content reaches the maximum irreducible water content, a very small
dry region is introduced in order to start the iteration. This was taken sufficiently larger than time step
∆t while small enough for the whole drying process, when the calculation results were influenced by
the choice of this initial, imposed dry region. After solving the dry region and wet region models at
time t, the distribution of water content, temperature, pressure, migration velocity of gas mixture, and
liquid and vapor diffusion velocity in the material at this time were obtained. Next, substituting the
vapor diffusion velocity and liquid velocity at the evaporation interface into Eq. (28), the position of
the evaporation front at the next time t + ∆t was computed, and then the next iteration starts.

Similarly, in order to improve the iteration accuracy, the position of the evaporation front can
also be determined using the idea of iterative correction, that is, the mass balance law can also be
checked in the calculation at time t using the following equation:

L yi(t) L H−yi(t)  L yi(t) 


.
∫ ∫ ϕ∆S(x, y)ρldxdy + ∫ ∫ ϕ∆Sb(x, y)ρldxdy =  ∫ 
∫ m(x,y)dxdy
0 0 0 0
 0 0

 LH−yi(t) 
 . L
 (32)
+ ∫ ∫ mb(x, y)dxdy + ∫ ρv,srvv,srdx ∆t
 0 0 0 
 

If the results did not satisfy the required accuracy for the above equation, we corrected the location
of the evaporation front calculated by Eq. (28) until convergence in terms of Eq. (32) was found.

4. Results and Discussion

Presemycki and Sturmillo studied the drying of brick [15]. In their experiment, the brick slab
had an area of 20 cm2 and was 5 cm thick. They also proposed a three-zone model made up of a wet
zone, an evaporation zone, and a dry zone. However, there was no information on how to determine
the position of the two interfaces and when the first and the second falling rate periods started.

The drying conditions and the physical properties of bricks are listed in Table 1. The
calculation results with the present model are shown in Figs. 3 to 6. From Figs. 3 and 4, it can be seen
that the variation of the average saturation and the temperature distributions during drying agree well
with the experimental results and are more accurate than the calculation results in Refs. 9–15. The
saturation distributions are shown in Fig. 5. The calculation results also show that the drying process
has three periods: a constant drying rate period, a first falling rate period, and the second falling rate
period. The first falling rate period starts at about 55 min, when the surface saturation reaches the
critical value 0.3 (Fig. 3). At about 125 min, the surface water content reaches the maximum
irreducible water content of the brick and the second falling rate period starts. At this time, the dry

Table 1. Physical Properties and Drying Conditions of Porous Material


ρ0 ϕ k Db,0 Ea/R H ρv,a Ta va T0 Sc Sir ηh ηm
1450 0.435 2.5 × 10−14 0.098 5200 0.05 0.0273 80 5 25 0.3 0.09 0.8 0.1

348
Fig. 3. Drying curve. " Experimental data; — Computed results by present model;
– – – Computed results in Ref. 15; - - - - Computed results by Darcy model.

region model and wet region model must be calculated simultaneously during the numerical simula-
tion. With an increase in drying time, the evaporation front receded continuously into the interior of
the materials and the dry zone extended gradually. Figure 3 also plots the drying curve, which is
obtained using Darcy’s model to calculate the movement of the gas phase. It shows that the movement
of the gas phase is weak and has little effect on drying behavior in the constant rate drying period and
the first falling rate drying period. However, as the drying proceeds and the dry zone expands, the gas

Fig. 4. Temperature distribution during drying. — Computed results by present model;


" Experimental data; – – – Computed results in Ref. 9.

349
Fig. 5. Saturation distribution during drying. — k = 2.5 × 10–14 m2; - - - k = 5.0 × 10–14 m2.

phase movement will be enhanced and becomes important for the drying behavior so that the
calculated error increases, using Darcy’s model in which the inertia and viscous motion of the gas
phase are neglected. The influence of the material’s permeability on drying behavior is shown in Fig.
6. For porous materials with higher permeability, the drying rate is slower in the constant rate drying
period and faster in the falling rate drying period. This is because the gradient of moisture is smaller
for these materials (Fig. 5), and the migration of moisture is slower in the constant rate drying period.
But in the falling rate drying period the drying rate is faster, because the movement of gas mixture
and the diffusion of vapor are easy for materials with high permeability.

Fig. 6. Influence of intrinsic permeability: (1) k = 2.5 × 10–14 m2; (2) k = 5.0 × 10–14 m2.

350
Literature Cited

1. Zhang Z. Investigation of heat and mass transfer in moist porous media, Ph.D. thesis, Huazhong
University of Science and Technology, 1994. (in Chinese)
2. Sherwood TK. The drying of solids. Ind Engng Chem 1929;21:12–16.
3. Sherwood TK. Application of the theoretical diffusion equation to the drying of solids. Trans
AIChE 1931;27:190–202.
4. King CJ. Freeze drying of foods. London: Butterworth; 1971.
5. Krischer O, Kast W. Die wissenschaftlichen grundlagen der tocknungstechnik. Berlin: Sprin-
ger; 1978.
6. Berger D, Pei DCT. Drying of hygroscopic capillary porous solids—a theoretical approach.
Int J Heat Mass Transfer 1973;16:293–302.
7. Eckert ERG, Pfender E. Heat and mass transfer in porous media with phase change. Heat
transfer 1978: Proc Sixth Int Heat Transfer Conf 1978;6:1–12.
8. Eckert ERG, Faghri M. A general analysis of moisture migration caused by the temperature
differences in an unsaturated porous medium. Int J Heat Mass Transfer 1980;23:1613–1623.
9. Chen P, Pei DCT. A mathematical model of drying process. Int J Heat Mass Transfer
1989;32:297–310.
10. Philip JR, De Vries DA. Moisture movement in porous materials under temperature gradients.
Trans Am Geophys Union 1957;38:222–232.
11. De Vries DA. Simultaneous transfer of heat and moisture in porous media. Trans Am Geophys
Union 1958;39:909–916.
12. Luikov AV. System of differential equation of heat and mass transfer in capillary-porous
bodies. Int J Heat Mass Transfer 1975;18:1–14.
13. Whitaker S. The transport equation for multi-phase system. Chem Eng Sci 1973; 28:139–147.
14. Whitaker S. Simultaneous heat, mass and momentum transfer in porous media: A theory of
drying. Advances in Heat Transfer 1977;13:119–203.
15. Przesmycki Z, Strumillo C. The mathematical modelling of drying process based on moisture
transfer mechanism. Drying ’85 1985;126–134.
16. Wang CY, Beckermann C. A two-phase mixture model of liquid-gas flow and heat transfer in
capillary porous media—I. formulation. Int J Heat Mass Transfer 1993;36:2747–2758.
17. Ilic M, Turner IW. Convective drying of a consolidated slab of wet porous material. Int J Heat
Mass Transfer 1989;32:2351–2362.
18. Nissan AH, Kaye WG, Bell JR. Mechanism of drying thick porous bodies during the falling
rate period. AIChE 1959;JI 5:103–110.
19. Chirife J. Fundamentals of the drying mechanism during air dehydration of foods. In:
Mujumdar AS, editor. Advances in drying. Washington, DC: Hemisphere; 1983. p 73–102.
20. Hsu CT, Cheng P. Thermal dispersion in porous media. Int J Heat Mass Transfer
1990;33:1587–1597.
21. Patanker S. Numerical heat transfer and fluid flow. New York: McGraw-Hill; 1980.

"F F F"

Originally published in Journal of Chemical Industry and Engineering, 48 (1), 1997, 52–59.
Translated by Zhe Zhang (postdoctoral fellow), Institute of Engineering Thermophysics, Chinese
Academy of Sciences, P.O. Box 2706, Beijing, P.R. China 100080.

351

You might also like