You are on page 1of 17

Chapitre 3: Modélisation sur Comsol

III.1 Numerical model of drying


III.1.1 Introduction
The drying model that we will present is based on the Finite Element Method (FEM) and the
calculations are performed on the COMSOL calculation software.

The model considered is a mesoscopic model with a representative volume element. Thus, it is very
difficult to predict accurately the different transfer modes at the microstructure scale. The simulation
of heat and water transfer at this scale is therefore not feasible. We therefore assume that the water
distribution within the material is homogeneous. Finally, we are only interested in the first drying
phase.

In the present work, we base ourselves first on experimental work in order to develop a numerical
model of thermo-hydric transfer. Then, in a second step, this model is validated through a
comparison of numerical results with experimental measurements obtained for an alumina sample.

III.1.2 Thermo-hydric behavior


In this section, we will present the equations of the models used such as the heat and mass (water
vapor) transfer model within an alumina paste exposed to hot air (Figure1).

The transport equations are respectively reacted by Fourier's law and Fick's law. The heat and mass
dT 2
=D t ×∇ T
transfer (Eq.1) are caused by a temperature gradient. dt
(1 )

where T is the temperature and D t is the thermal diffusion coefficient which is calculated as a
function of the thermal conductivity λ, the density ρ and the heat capacity of the material Cp.
According to Fick's law (Eq 2), the mass flux is proportional to the moisture concentration gradient:
dC 2
=D w × ∇ C
dt
( 2)

where C is the water concentration and D w is the mass diffusion coefficient.


Figure 1 Geometry of the drying model and its assumptions [14]
The thermo-hydric model allows to represent the heat and mass transfer phenomena. It offers a
good accuracy of the results compared to the experimental results with a limited time of resolution.

The heat and mass transfer equations are solved simultaneously. The coupling between these
equations is ensured by the two following boundary conditions (Eqs.3 and 4):
ɸT =D w Lw ∇ C M w +h ( T a +T s )
( 3)

ɸC =k ¿
¿

where ɸT is the surface heat flux, Lw is the latent heat of evaporation and sorption, M w s the molar
mass of water, R is the perfect gas constant, and h is the heat transfer coefficient.

In Equation 4, ɸC is the transfer flux, k is the mass transfer coefficient and C a et C lim ¿¿ are the
concentration of water in the air and in the boundary layer, respectively. These are determined as
follows:

P v ,sat
C a=RH ×
RTa
( 5)

C Pa (6 )
lim ¿=a w × ¿
RT s

T a is the air temperature and et T s is the temperature of the solid. a w is the water activity which is
defined as the relative humidity within the material.

Pv
a w=
P v ,sat
(7 )

It can be determined experimentally. A study in [14] [13] showed that water activity can be related to
[ ( )]
B −1
A
a w = 1+
the water content X. X where A and B are two parameters to be identified
( 8)
experimentally.

The scientific challenge of this work is the coupling of the thermal and moisture terms.

The heat and mass transfer problems are solved to describe the exchange phenomena and
determine the temperature and moisture content distribution of the part to be studied. This
resolution is made independently of the mechanical resolution. The numerical model uses 2
calculation modules implemented in COMSOL:

 A heat transfer module to solve the thermal problem.

 A dilute species transfer model to model mass transfer.

Furthermore, this model requires certain physical properties and parameters in the form of
inputs that depend on the water content (X).
Thus, in the example of alumina, the water activity a w can be described according to the
Oswin model where the parameters A and B are determined experimentally (Figure 2) as a
function of the water content [19]

Figure 2: OSWIN water activity fitting for alumina at 20°C


(A=0.262, B=2.6623)

Note that the water content (X) is related to the water concentration by the following expression:
C × Mw
X=
ρsec
( 9)
- Mass diffusion coefficient D w

Following the study of the evolution of the water mass diffusion coefficient D w as a function of the
drying kinetics, we can distinguish 2 periods. In the first period, only water and solid material coexist.
The water distribution is then uniform and a maximum value is associated with D w . Using the same
sample size and drying conditions, we then reduce the diffusion until a water concentration gradient
appears. The last value that leads to a uniform water concentration was used in the mode (
−5 2 −1
6. 10 m . s ). In the second phase where a concentration gradient appears, we introduce a
diffusion coefficient that takes into account the variation of the water content. In the example of
alumina the relationship D w and water content (X) is written [19]:
−9 0.3 X
D 2=2 ×10 e
( 10 )

The transition between the two drying periods is expressed by the following relationship:

Dw ( X )= D2+ erf
( X −X f , p
0.5
2 dX )( D1−D 2)

( 11 )

X f , p is the water content for which the first drying period ends, dX is the standard deviation and
erf is the error function:

x
2
∗∫ exp (−t ) dt
2
erf =
√π 0
(12 )

The evolution of the diffusion coefficient in the example of alumina is presented in Figure 3:

Figure 3: Fittage of the effective water diffusion coefficient as a function of water content
- Thermal conductivity λ

The thermal conductivity is not a priori constant and depends strongly on the water content (X). In
the example of alumina [16], we find three distinct regimes. From a practical point of view, the
experimental points are fitted (Figure 4) according to the following linear equation:

λ=aX + b
( 13 )

where a and b are constants to be determined.

The increase in thermal conductivity λ is explained by the increase in water evaporation within the
material. This is because of the volume fraction of the solid has a higher conductivity than water.
Then, the decrease in thermal conductivity is explained by the fact that water will be replaced by air
that has a thermal conductivity lower than that of water.

Figure 4: Fittage of thermal conductivity as a function of


water content

The numerical study of the drying process (or humidification) requires a number of input

parameters. For the case of alumina, Table.1 below summarizes the constants and parameters

to be introduced [19]:
Parame`tres Symbole Valeur Unite´

Density of water ρw 1000 kgm 3

Density of air ρair 1.2 kgm 3

Density of alumina ρsolide 3986 kgm 3
Pore volume fraction νp 0.5

Apparent density of dry matter ρsec ρsolide (1 − νp) kgm 3
− −
Apparent density of the dry matter for a R 8.314 JMol 1K 1
given porosity

Constant of perfect gases Mw 0.018 kgmol 1
Cpsolide −
Molar mass of water 780 Jkg 1 −1
− K
Heat capacity of alumina Cpw 4180 Jkg 1 −1
Cpair − K
Thermal capacity of water 1000 Jkg 1 −1
K
Heat capacity of air Lef 1
−1
Lewis coefficient Lw(T) 1000 (2501-2.43(T-273.15)) Jkg
Latent heat of vaporization of water Ps(T) 133.22 exp[46.784-(6435/T)-3.868 ln(T)] Pa

III.2 Validation of the numerical model


In the previous section, we presented an approach allowing the resolution in the transient domain
of a thermo-hydric problem of a material sample subjected to thermal and hydric gradients. We
now consider the experimental work done on alumina as an application framework. Different
configurations are studied in order to validate the numerical model.

The proposed problem is that of convective drying of a cubic shaped alumina sample of side L =
20mm or the bottom face exchanges neither heat nor mass with the outside as presented in Figure
5. The part is placed on a furnace with a constant temperature T=40°C and a humidity of 29%. It is
subjected to humidity gradients for an initial temperature T init=20° C . The initial water content is
30%. Figure III.6 below shows the state of the material after two and a half hours of drying.

Figure 5: Initial conditions imposed on a cubic block of alumina deposited in a furnace


at 40°C

From a hygroscopic point of view, the lower part contains more water than the rest of the domain.
We can clearly see the effect of thermal and hygroscopic exchanges with the air. This validates in a
qualitative way our model. From a quantitative point of view, the numerical model allows to
access several physical quantities (Figure 7) and this at any point of the domain. For example, we
can follow the water content in the sample. We can also measure the variation of the temperature
as a function of time.

Figure 6: Variation of the water concentration field and temperature represented in 3D

Figure 7: Variation of water content and temperature with time

A comparison is made between experimental and numerical results concerning temperature and
moisture content for drying at 40°C. The first phase, at a constant drying rate, is characterized by a
constant drying rate per unit area clearly identified on the water content curve during the first 4-5
hours. Indeed, the surface exposed to the drying of the sample behaves as a free water surface.
During this phase, the heat flow and the latent heat of the evaporation water tend to establish an
equilibrium state characterized by a constant temperature of the sample surface (≈ 27°C). Before this
is established, a transient regime is observed which can be explained by the initial sample
preparation temperature (20°C) which is lower than the oven temperature (40°C). Then, when the
drying rate is slowed down by disturbing the equilibrium of the evaporation surface, a strong
increase of T is observed. The model shows a good agreement with the experimental data, either in
terms of moisture content or in terms of surface temperature. On the other hand, the small
discrepancy between the experimental and numerical results cannot be ignored. This may be related
to the experimental preparation of the sample at room temperature (20°C) before drying it at 40°C
or to the choice of the heat transfer coefficient h.

Conclusion
In this chapter, a three-dimensional macroscopic numerical model has been developed and validated
in the context of the study of alumina. This model is based on a coupling of the heat and mass
equations, but its main advantage lies in the fact that changes in the determining physical properties
with water content are taken into account. The solution of these equations was performed using
FEM.
Chapter 4: Application of the Numerical Model

IV.1 Introduction
In chapter III, we have seen that following the introduction of a set of thermophysical and
hydric properties in a numerical model, results can be obtained with an acceptable degree of
realism. Our objective is now to apply our numerical model to the wetting of candidate
materials (CEM1 cement, gypsum plaster) considered to perform SAW sensor operation
tests.

We consider Portland CEMI cement (sample C5) [15] whose input parameters are given in
Table 2 and whose isothermal sorption curve is given in Figure IV.1. The data of the sorption
curve of sample C5 have been entered into the COMSOL software.

Paramètres Valeur
Porosity 14.8 %
Mass diffusion coefficient D w 10
−10
m²/s
Heat transfer coefficient h 40 W/m². K
Saturation vapor pressure of water Ps(T) = exp (13.7 − (5120/T)) : Pa
Thermal conductivity λ 1.5 W/m. K

Table 2: Input parameters of the numerical model

Figure 8: Variability of sorption isotherms for different samples at 20°C (focus is on sample
C5) [15]
IV.2 Numerical model of water vapor sorption in cement
The proposed problem is that of the sorption of a dry cement sample of cubic shape with side L
whose admissible value depends on the dimensions of the sensor and varies between 0.02m and
0.1m. All the faces of the cube exchange heat and humidity with the outside. The room is placed in a
climate chamber at constant temperature T= 20°C as shown in Figure 9 and is subjected to time-
varying relative humidity gradients (Eq.14) at an initial temperature T init . This choice for the relative
humidity is to avoid the appearance of cracks in the sample

{
HR=0.1 ,∧si(t < X 1 )
HR=0.4 ,∧si (X 1 ≤ t < X 2 )
HR=f ( t )=
HR=0.8 ,∧si(X 2 ≤t ≤ X 3)
HR=1 ,∧si(X 3 >t )
( 14 )

Figure 9: Conditions imposed on the CEM1 cement sample during the sorption
process

Results and discussion:

Porous materials such as cement have the ability to exchange moisture with the surrounding air.
The increase in the humidity of the air in the vicinity of the material leads to an increase in its
apparent mass. A fixation of water molecules on the surface of the pores appears, it is the
−1 −1
phenomenon of adsorption. Liquid water has a thermal conductivity ( λ air =0.6 W . m . K ) thirty
−1 −1
times higher than that of dry air ( λ air =0.6 W . m . K ). Thus the presence of water in the porous
medium will modify the overall thermal conductivity of materials.
Figures 10, 11 show the impact of block dimensions on the time required to reach a state of thermo-
hydric equilibrium where a w =H Ra. Table 3 summarizes the time needed to reach the different
levels of relative humidity-imposed HR=0.1, 0.4, 0.8, and 1, we see that for L=0.02m we need 9 days
to reach the 10% of the relative humidity of the air imposed. Note that for L=0.1m the process
requires 200 days (7 months) to reach the same value of relative humidity. On the other hand, each
time the size of the sample increases the time required to go from one stage of thermo-hydric
equilibrium to another. This is due to the increase in the volume of the sample which requires more
time for water to diffuse from the surface to the interior of the concrete sample.

From a theoretical point of view, numerical simulations have shown that the sorption time is very
sensitive to the dimensions of the sample and to the imposed values of the relative humidity of the
air.

In order to meet the objectives of the project, the CEM1 cement, although close to the refractory
concrete, turns out to be a bad practical choice because it requires much too important sorption
times (up to 28 months L=0.1m).

Figure 11: Variation of water content for different


dimensions as a function of time

In the following section, we consider another material, gypsum plaster, as a study material. This
material presents indeed an important rate of porosity as well as a coefficient of diffusion more
important than the cement, parameters which influence greatly the time of sorption.
IV.3 Numerical model of water vapor sorption in gypsum plaster
We apply the same approach followed during the study of sorption of cement CEM1 for gypsum
plaster. Thus, we obtain, for the same steps as seen previously, the variation of the water content
and water activity as a function of time. Experimental data from the literature [16] [17] for gypsum
plaster are presented in Table 4:
Sorption isotherms of gypsum plaster from the work of MATIASOVSKY and TAKACSOVA. The
sorption curve used in the following work is presented in Figure 12.

Paramètres Valeur
Porosity 55 %
Mass diffusion coefficient D w 2×10−8 m²/s
Heat transfer coefficient h 40 W/m². K
Pressure of the saturated vapor of water Ps(T) = 101325*exp (13.7 − (5120/T)) : Pa
Thermal conductivity λ 0.48 W/m.K

Table 4: Input parameters of the numerical model

Figure 12 : Isothermes de sorption pour le plâtre [18]

The proposed problem is that of the sorption of a sample of gypsum plaster of cubic shape
of side L which varies from 0.02m to 0.1m. All the faces of the cube exchange heat and
mass with the external environment. The specimen is placed in a climate chamber at
constant temperature T= 20°C as shown in Figure 13 and subjected to time-varying relative
humidity gradients (Eq 15) at an initial temperature T init =20°C:

{
HR=0.1 ,∧si(t < X 1 )
HR=0.4 ,∧si (X 1 ≤ t < X 2 )
HR=f ( t )= HR=0.8 ,∧si(X 2 ≤t ≤ X 3) ¿
HR=1 ,∧si ( X 3 >t )
¿
( 15 )

Figure 13: Conditions imposed on the gypsum plaster sample during the
process

Comparison of results between CEM1 cement and gypsum plaster

Figures 14, 15 show again that the time needed to reach the RH set points for gypsum varies with the
size of the sample. According to Tables 5, 6 the time needed to reach the set values of relative
humidity for gypsum (maximum of 5 days for L=0.1m) is much shorter than for concrete (maximum
28 months for L=0.1m). This difference is mainly due to the high value of the porosity for gypsum 55
% against 14.8 % for cement and to the value of the mass diffusion coefficient (2. 10−8 m²/s against
−8
10 m²/s for cement).
Thus, in order to validate the performance of the sensor, it is preferable to work with gypsum, which
has a shorter and more efficient test time than cement CEM1.

Figure 14: Variation of water activity for different dimensions


as a function of time

HR=0.1 HR=0.4 HR=0.8 HR=1


L=0.02 m t ≤ 1.2h 1.2h ≤ t ≤ 2.2h 2.2h ≤ t ≤ 3.2h t > 3.2h
L=0.04 m t ≤ 4.5h 4.5h ≤ t ≤ 8.5h 8.5h ≤ t ≤ t > 12.5h
12.5h
L=0.06 m t ≤ 10h 10h ≤ t ≤ 20h 20h ≤ t ≤ 28h t > 28h
L=0.08 m t ≤ 18h 18h ≤ t ≤ 34h 34h ≤ t ≤ 47h t > 47h
L=0.1 m t ≤ 20h 28h ≤ t ≤ 52h 52h ≤ t ≤ 72h t > 72h

Table 5: Time required to reach the different target values of relative humidity depending on the size of the cement block

Figure 15: Variation of water content for different dimensions


as a function of time
HR = 0.1 HR = 0.4 HR = 0.8 HR = 1
L = 0.02M Be 210h 240h 200h 350h
´ton
Plaˆtr 1.2h 1h 1h 2.8h
e
L=0.1m Be 4800h 6200h 5000h 4000h
´ton
Table 6: Plaˆtr 20h 24h 20h 48h Comparison
Table e

Conclusion

The examination of numerous thermo-hydric transfer configurations with different types of materials
(alumina, CEM1 cement, gypsum plaster) allowed us to conclude that our model is capable of
predicting some useful study parameters such as the determination of the time required to validate
the operation of the moisture and pressure sensors embedded in the material.
Conclusions Générales

The work carried out during this internship led to the implementation of a numerical tool
contributing to the study and the validation of the operation of innovative sensors for the in-situ
monitoring of refractory concrete during the drying process.

First of all, in chapter II, a detailed study of the drying phenomenon was presented. Such a process is
indeed quite complex due to the simultaneous physical phenomena of heat and mass transfer. Our
challenge consists in setting up a reliable and powerful numerical tool making it possible to predict
via a set of input data that involves the evolution of the drying process whatever the severity of the
temperature and moisture conditions. In order to meet this need, a bibliographical study was carried
out in order to highlight the experimental data necessary for the numerical simulation. These data
include water content, water activity, relative humidity, specific heat, thermal conductivity, and
water diffusion coefficient.

In the second step, a three-dimensional numerical model has been developed in chapter III dedicated
to the drying of porous materials. This model is based on a coupling of the heat and mass equations
and on the boundary layer theory. The resolution of these equations is based on FEM and the
calculations are performed using the COMSOL calculation software. The validation of this model was
performed in several configurations (with different initial conditions and sample geometries) by
comparison with experimental results from the study of convective drying of an alumina paste.

Finally, the developed numerical model was applied in chapter IV to the water vapor sorption within
two porous materials: cement CEM1 and gypsum plaster, two materials considered to perform tests
of the sensor operation. According to the results obtained, gypsum plaster better meets the needs of
the project in terms of thermo-hydric performance and test duration. It is a good candidate to
validate the performance of the sensor in terms of test time by comparing it with concrete.

The model can play a dual role:

- Allows to know which material is suitable for the sensor, by comparing the data of each candidate
material (moisture, temperature...).

- It can also play the role of the sensor itself as it can give temperature and water content curves as a
function of time after drying or wetting (it can predict the drying state of the concrete)

=> From the result of this work we have a model that gives us curves of temperature and water
concentration that reflects the state of any material after drying or wetting.
Bibliography

[19] Siham OUMMADI. Drying behaviour of ceramic green bodies: experimental characterization
and numerical modeling. Theses , 2019.

You might also like