Subido por William Díaz

Mathematical approaches for use of analytical solutions in (2)

Anuncio
Journal of Food Engineering 68 (2005) 233–238
www.elsevier.com/locate/jfoodeng
Mathematical approaches for use of analytical solutions in
experimental determination of heat and mass transfer parameters
Ferruh Erdoğdu
*
Department of Food Engineering, (Gida Muh. Bol.), University of Mersin, Ciftlikkoy-Mersin, Mersin 33343, Turkey
Received 2 January 2004; accepted 31 May 2004
Abstract
Heat (heat transfer coefficient and thermal diffusivity values) and mass (mass transfer coefficient and diffusion coefficient values)
transfer parameters are crucially important for characterization and modeling of food processing operations. Since they are important function and property of interested material and medium itself, experimental determination of these values would be valuable.
Analytical solutions for regular geometries (infinite slab, infinite cylinder and sphere) have a broad use in experimentally determining these parameters. These solutions, with use of experimental data, might give a greater advantage over use of other methods, e.g.
the lumped system approach or empirical equations. Effective use of numerical solution techniques with experimental data would
enable simultaneous determination of the related parameters with the experimental data. The experimental data and the analytical
solutions may also be used to determine the thermocouple locations to use in the model validation studies. This technique would be
much faster and easier compared to the other methods used for this objective.
2004 Elsevier Ltd. All rights reserved.
Keywords: Analytical solutions; Heat and mass transfer parameters
1. Introduction
Heat (heat transfer coefficient and thermal diffusivity)
and mass (mass transfer and diffusion coefficients) are
required parameters to model heat and mass transfer
processes in food processing operations. Thermal diffusivity ða ¼ qck p Þ and diffusion coefficient (D) are the material properties, and especially the composition of the
product and process temperature might affect these
properties. On the other hand, the heat (h) and mass
(kc) transfer coefficients depend on the thermo-physical
properties of the medium, characteristics of the product (size, shape, surface temperature and surface
roughness), characteristics of fluid flow (velocity and
*
Tel.: +90 5338120686; fax: +90 3243610032.
E-mail addresses: [email protected], ferruherdogdu@
yahoo.com
0260-8774/$ - see front matter 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jfoodeng.2004.05.038
turbulence) and heat and mass transfer equipment
(Rahman, 1995). Even though, there have been numerous expressions in the literature to determine these properties (different Nusselt and Sherwood number
correlations as a function of Reynolds or Grashof and
Prandtl numbers), experimental determinations to
reflect the process parameters are important. Another
possibility is to use the experimental time–temperature
or time–mass concentration data to determine these
parameters. As an experimental approach in heat transfer studies, lumped system method is one preferred
method to determine the heat transfer coefficient even
though it does not reflect the real process due to the
use of a highly conductive material (e.g., copper, aluminum, etc.). Analytical solutions for regularly shaped
geometries (infinite cylinder, infinite slab and sphere)
with proper initial and boundary conditions with experimentally obtained data from food material itself may
be used to experimentally determine these properties.
234
F. Erdoğdu / Journal of Food Engineering 68 (2005) 233–238
Nomenclature
A
a
cp
D
E
n
h
J0, J 1
k
kc
L
k
m
l
constant
thermal diffusivity, m2/s
specific heat, J/kg K
diffusion coefficient, m2/s
experimental data
characteristic length, m
heat transfer coefficient, W/m2 K
the first kind 0th and 1st order Bessel
functions
thermal conductivity, W/m K
mass transfer coefficient, m/s
half-thickness for an infinite slab, m
root of Eqs. (5)–(7)
slopes of temperature or concentration ratio
vs time curves, 1/s
dynamic viscosity, kg/m s
In fact, proper use of analytical solutions with the
experimentally obtained data might give a greater
advantage over the use of other methods, e.g. lumped
system approach or empirical equations. The analytical
solutions with the experimental data may also be used to
find out the precise geometric location of the thermocouples placed in the samples to measure the temperature change. In fact, the precise thermocouple location
is crucially important in model validation studies rather
than in determining the heat and mass transfer parameters. Therefore, the objective of this study was to explain
the methodologies to use analytical solutions of regularly shaped geometries with the experimentally obtained data to determine the heat and mass transfer
parameters and thermocouple locations where the temperature measurement was taken.
g
X
/
/i
/1
w
w
r, x
R
q
m
heat or mass transfer coefficient in heat or
mass transfer cases
thermal diffusivity or diffusion coefficient,
m2/s
temperature or mass concentration
initial temperature or mass concentration
medium temperature or mass concentration
dimensionless
temperature
or
mass
concentration ratio
volume average dimensionless temperature or
mass concentration ratio
distance from the center, m
radius for an infinite cylinder or a sphere, m
density, kg/m3
kinematic viscosity, m2/s
cylinder, and n = 2 for sphere); and X is thermal or mass
diffusivity (m2/s).
Solutions for Eq. (1), for initial condition of uniform
temperature or mass distribution and boundary conditions of central symmetry and convective boundary at
the surface, are given for these geometries as follows
(Carslaw & Jaeger, 1986).
Infinite slab:
/ /1
/i /1
1 X
2 sin kn
x
2
¼
cos kn expðkn FoÞ
L
kn þ sin kn cos kn
n¼1
w¼
ð2Þ
Infinite cylinder:
/ /1
/i /1
1 X
2
J 1 ðkn Þ
x
2
¼
2
k
FoÞ
expðk
J
0
n
n
kn J 0 ðkn Þ þ J 21 ðkn Þ
L
n¼1
w¼
2. Mathematical background
Governing differential equations and solutions for
infinite geometries (infinite slab, infinite cylinder and
sphere) are given as follows with the mostly encountered
initial (uniform distribution of temperature or mass concentration) and boundary condition (surface convection).
Governing differential equation:
1 o
1 o/
n o/
x ð1Þ
¼ xn ox
ox
X ot
where / is temperature or mass concentration; x is location in geometry to characterize temperature or mass
concentration (m), distance from the center; n is a characteristic number (n = 0 for infinite slab, n = 1 for infinite
ð3Þ
Sphere:
/ /1
/i /1
1 X
2 ðsinkn kn coskn Þ sinðkn Rr Þ
2
¼
expðkn FoÞ
kn Rr
kn sinkn coskn
n¼1
w¼
Xt
n2
ð4Þ
where Fo Fo ¼
is Fourier number (n is L, halfthickness for an infinite slab and R, radius for an infinite
cylinder or a sphere; X is thermal diffusivity or diffusion
F. Erdoğdu / Journal of Food Engineering 68 (2005) 233–238
coefficient), and the knÕs are the roots of Eqs. (5)–(7) for
infinite slab, infinite cylinder and sphere, respectively:
Bi ¼ k tan k
ð5Þ
J 1 ðkÞ
J 0 ðkÞ
ð6Þ
Bi ¼ k k
Bi ¼ 1 ð7Þ
tan k
where Bi Bi ¼ gn
is Biot number (g is heat or mass
k
transfer coefficient, and k is thermal conductivity or X,
diffusion coefficient in a mass transfer case), and J0
and J1 are the first kind 0th and 1st order Bessel functions.
In heat transfer related calculations, use of Eqs.
(2)–(4) would be enough since the temperature change
at a certain location inside the geometry is needed or
experimentally obtained. However, for mass transfer related calculations, these equations should be integrated
through the whole volume since the mass transfer experimental results would generally be obtained through the
whole volume rather than at a certain location. Integration of Eqs. (2)–(4) through the whole volume results in
Eqs. (8)–(10) for infinite slab, infinite cylinder, and
sphere, respectively:
Infinite slab:
/ /1
/i /1
1 X
2
sin2 kn
2
expðkn FoÞ
¼
kn ðkn þ sin kn cos kn Þ
n¼1
w¼
diffusion coefficient is constant, this first term approach
may be easily used to determine these parameters with
the known heat or mass transfer coefficient value.
3. Methods
3.1. Determination of constant thermal diffusivity or
diffusion coefficient value
Let us assume that the temperature change at a certain location of a spherical object in a highly agitated
medium was recorded to determine the thermal diffusivity. Due to the agitation, heat transfer coefficient and
hence the Biot number may be assumed to be infinite.
Then, roots of Eq. (7) may be determined to be
(p,
3p, . . . and so on). The roots would be
p 2p,
3p 5p
;
;
; . . . for an infinite slab or (2.4048, 5.5200,
2 2
2
8.6537, . . .) for an infinite cylinder when the Biot number
is infinite. Due to the fact that the temperature ratio
//1
becomes linear after a certain time (Fo > 0.2),
/i /1
the first term of Eq. (4) in case of sphere is then used
to characterize the linear change in that region. When
the natural logarithm of both sides of Eq. (4) is taken
with the first term approximation (n = 1), following
equation is obtained (k1 = p):
/ /1
k2 X
p2 X
t
ð11Þ
ln
¼ A1 1 2 t ¼ A 1 /i /1
R
R2
where,
ð8Þ
Infinite cylinder:
/ /1
w¼
/i /1
"
#
1
X
4
J 21 ðkn Þ
2
expðkn FoÞ
¼
2
2
2
n¼1 kn J 0 ðkn Þ þ J 1 ðkn Þ
235
ð9Þ
Sphere:
/ /1
/i /1
"
#
2
1
X
6 ðsin kn kn cos kn Þ
2
¼
expðkn FoÞ
3
kn sin kn cos kn
n¼1 kn
w¼
ð10Þ
As seen in Eqs. (2)–(4) and (8)–(10), it is important to
know how many terms of the infinite series solutions are
required to obtain a correct solution. It is a general
knowledge that use of the first term would be enough
when the Fourier number is greater than 0.2 since the
temperature or mass ratio change after that certain time
would be linear. As long as the thermal diffusivity or the
2 ðsin k1 k1 cos k1 Þ sin k1 Rr
A1 ¼ ln
k1 sin k1 cos k1
k1 Rr
r sin p R
¼ ln 2 p Rr
ð12Þ
As seen in Eq. (11), slope (m) of the temperature ratio vs
2
k2 X
time curve would be equal to m ¼ R1 2 ¼ pRX
2 . Then,
with the known slope and radius of the spherical material, thermal diffusivity value may be determined. As
one realizes, this method does not require the knowledge
of the location in the material where the experimental
data is recorded since, regardless of the location, temperature ratio curves have the same slope after the linearization starts (Chau, 1995–2000). Eq. (11) also shows that
the slope is independent of the location. Fig. 1 shows
the change in simulated temperature ratios obtained
using the analytical solution at different locations
(Rr ¼ 0, Rr ¼ 0:5 and Rr ¼ 0:97) of a spherical object (30
mm in radius) cooked in agitated boiling water. Rr ¼ 0
is the center while Rr ¼ 0:97 is the location just underneath
the surface. As seen in Fig. 1, temperature ratios become
linear after a certain time (e.g. t > 1000 s for the center),
and slopes are equal for each location. The same
approach may also be easily applied to determine the
236
F. Erdoğdu / Journal of Food Engineering 68 (2005) 233–238
ln(Temperature Ratio)
0
-1
r/R=0
r/R=0.5
-2
r/R=0.97
-3
-4
-5
-6
-7
-8
0
500
1000
1500 2000
Time (s)
2500
3000
3500
Slopes of the linear parts for each curve are: - 0.0013 1/s
Fig. 1. Simulated temperature ratios taken in different locations of a
spherical object.
diffusion coefficient during mass transfer phenomena
(e.g., diffusion coefficient of moisture during drying) via
the use of Eq. (10) when the changes in volume are
known.
As explained, it is very easy to determine the thermal
diffusivity or diffusion coefficient as long as the experimental data is available in the range of high Fourier
number (Fo > 0.2). For the heat transfer case, it is relatively easy to obtain the data in this range since the thermal diffusivity values are much higher than the diffusion
coefficient values. For example, thermal diffusivity value
of unfrozen red meat is in the range of 1.2–1.3 · 107
m2/s (Rahman, 1995) while the diffusion coefficient
value; e.g. for sodium tripolyphosphates (prepared as
2% solution, w/v) in the red meat, is in the range of
1011 m2/s. The required times for Fo > 0.2 for red meat
of 20 mm in thickness (infinite slab shaped) may be determined as 2.5 min for the heat transfer case and more than
500 h for mass transfer case (Ünal, Erdoğdu, Ekiz, &
Özdemir, 2004). In the latter case, it would not seem to
be possible to determine the diffusion coefficient value
from the slope of the concentration ratio vs time curves.
In this case an alternative approach must be used.
Let us assume that the concentration of sodium tripolyphosphates were determined experimentally in red
meat while the meat sample is dipped in the solution
for 30 min. As explained, it is not possible to get a higher
Fourier number in this case. To solve this problem, Eq.
(8) can be written as:
/ /1
/i /1
2
2
2 Xt
¼ A1 exp k1 2 þ A2
k1
k
n
2
X
t
exp k22 2 þ n
w¼
for n ¼ 1; 2; . . .
"
1
X
2 kn t
2 Xt
0
An exp kn 2
f ðXÞ ¼ n2
n
n¼1
Xnþ1 ¼ Xn f ðXn Þ
f 0 ðXn Þ
#
ð16Þ
ð17Þ
In this method, just knowing the concentration ratio (or
temperature ratio in the heat transfer case) at any time
would be enough to determine the diffusion coefficient
value instead of using a set of experimental data. The
constant diffusion coefficient value assumption may also
be checked with this procedure using the experimental
data obtained at different times.
3.2. Determination of variable thermal diffusivity or
diffusion coefficient value
The variable thermal diffusivity is not common. However, Ünal et al. (2004) reported variable diffusion coefficient values in the case of sodium tripolyphosphate
diffusion in red meats. The variability of the diffusion
coefficient may be tested using the procedure explained
above (Eqs. (15)–(17)). If the variation continues
through the whole process, it may be easier to determine
an averaged value to describe the whole process, or
numerical finite difference techniques should be used to
take the variability into account. In this case minimization of sum of squares (S) of the differences between the
experimental data and results of the analytical solutions
may be used to minimize Eq. (18) as follows:
S¼
n
X
2
ðw EÞi
ð18Þ
i¼1
ð13Þ
where A values are given by:
sin2 kn
An ¼
kn þ sin kn cos kn
In Eq. (13), the only unknown is the diffusion coefficient
value (X) assuming the mass transfer coefficient of the
medium, therefore the Biot number and the k values
through Eq. (5), is known. In this case Eq. (13) should
be solved to determine the X value. Newton–Raphson
or any other numerical technique may be applied to
solve this equation. When the Newton–Raphson method is applied, the iterative solution of Eqs. (15)–(17) give
the result for X value.
1 X
2
Xt
An exp k2n 2
ð15Þ
w¼0
f ðXÞ ¼
kn
n
n¼1
ð14Þ
where n is the number of experimentally obtained data,
and E is the experimental data. To determine the diffuoS
sion coefficient value, to minimize S, oX
should be 0.
When this is applied to Eq. (18), Eqs. (19) and (20)
are obtained:
n oS X
ow
¼
2 ðw EÞ ¼0
oX
oX i
i¼1
ð19Þ
F. Erdoğdu / Journal of Food Engineering 68 (2005) 233–238
237
n
i
o
n h
oS X
2
þ k22 A2 exp k22 Xt
þ E 2kn21 t A1 exp k21 Xt
2kn22 t A2 exp k22 Xt
A1 exp k21 Xt
¼0
2
¼
2
2
2
2
k1
n
n
n
n
i
oX i¼1
i
ð20Þ
oS
From Eq. (20), it would be obvious that for oX
to be 0:
n
X
2 k1 t
Xt
A1 exp k21 2
2
n
n
i¼1
2 k1 t
2 Xt
A2 exp k1 2 ¼ 0 ð21Þ
n2
n
i
and Eq. (21) may be easily solved using Newton–
Raphson or any other numerical solution technique
for X value.
3.3. Determination of constant heat or mass transfer
coefficient
Heat or mass transfer coefficient may also be determined using the analytical solutions when a thermal diffusivity- or diffusion coefficient-known material is used.
In this case, using the slope (m) of the temperature
ratio
pffiffiffiffiffiffiffi
vs time curve, k1 is determined (k1 ¼ n mXÞ, and it is
then used to determine the Biot number (Bi) and therefore the heat or mass transfer coefficient. Of course, in
addition to the thermal diffusivity, thermal conductivity
of the experimental material must also be known in the
heat transfer case. In spite of the easy use of this method,
general approach in the literature is to assume an infinite
heat or mass transfer coefficient and then to determine
the thermal diffusivity or diffusion coefficient values.
However, just knowing heat transfer coefficient may also
enable to determine mass transfer coefficient and diffusion coefficients through the so-called analogy of
Chilton–Colburn when a simultaneous heat and mass
transfer is occurring (e.g. during drying process):
kc ¼
23
h
Pr
q cp
Sc
ð22Þ
where kc is mass transfer coefficient (m/s), h is heat
transfer coefficient (W/m2 K), q (kg/m3) and cp (J/kg K)
are density and heat capacity of the heating medium,
c l
respectively, Pr ðPr ¼ pk Þ is Prandtl number where l
is dynamic viscosity (kg/m s), and Sc ðSc ¼ Dm Þ is Schmidt
number where m is kinematic viscosity (m2/s) and D is
diffusion coefficient (m2/s). Basically, knowing mass
transfer coefficient with experimental data to determine
the diffusion coefficient leads to Eqs. (23) and (24).
Then, solving this equation leads to the determination
of k value (Eq. (23)) and therefore the diffusion coefficient value through the slope of the concentration ratio
vs time curve (Eq. (24)).
k1
kc R
¼
Bi ¼ 1 tan k1
m R2 =k21
m R2
¼X
k21
ð23Þ
ð24Þ
where R is the characteristic dimension (radius for
sphere). The solution to determine the kc value is not
as easy as it seems to be since the D value should also
be known through the Sc (Schmidt) number. Therefore,
assuming the D value is known, through the average
temperature of surface and medium, using the following
steps lead to determine the X value:
1. Determine kc value from Eq. (22),
2. Determine k1 value from Eq. (23) (Newton–Raphson
or any other method for a nonlinear equation may be
used), and
3. Determine X value from Eq. (24).
As long as the Chilton–Colburn analogy may be used,
this methodology should work without any problem.
However, one should not forget that there are limitations on the use of heat-mass convection analogies.
Heat-mass convection analogies are generally valid for
low mass flux cases so that mass transfer does not affect
the surface flow, and heat transfer correlations are
developed for smooth and constant surface temperature
situations. Therefore, an extra pre-caution should be
taken to find out the given conditions before applying
the heat-mass analogies for more accurate results.
3.4. Determination of thermocouple locations
In the heat transfer validation studies, the knowledge
of thermocouple locations, where the experimental data,
is extremely important to correlate the simulation results
with the experimental data. Since it is not easy to find
out where the tip of the thermocouple, it is a common
way to take an X-ray of the material when the thermocouple is still in it and then determine the location
(Anderson & Singh, 2002). The other way might be to
cut thin slices from the material until the tip was caught
(Erdogdu, Balaban, & Chau, 1998). Both these ways are
time consuming. The analytical solutions may be an
alternative approach to determine the thermocouple locations. If we go back to Fig. 1 again for this purpose,
the intercepts of the three temperature ratios may be
238
F. Erdoğdu / Journal of Food Engineering 68 (2005) 233–238
determined as 0.6821, 0.2416, and 2.6727 for Rr ¼ 0,
r
¼ 0:5 and Rr ¼ 0:97, respectively. Now, let us try to
R
determine the thermocouple locations ðRr Þ values from
these data. Then, using Eq. (11) and applying linear
regression to the linear part of the data points, the following relationship is obtained for the second point
(Rr ¼ 0:5):
sin p Rr
A1 ¼ ln 2 ¼ 0:2416
ð25Þ
p Rr
Now, the objective is to determine the Rr value from
this equation. When Newton–Raphson method was applied, the following equations are obtained to solve:
r
r
r
ð26Þ
f
¼ 2 sin p p expð0:2416Þ
R
R
R
r
r
f0
¼ 2 p cos p p expð0:2416Þ
ð27Þ
R
R
Then, the iterative solution of Eqs. (26) and (27), starting with the initial guess of 1, leads to find Rr value as 0.5.
proaches, it is required that one of the parameters
should be known to determine the other: heat or mass
transfer coefficient should be known to determine the
thermal diffusivity or the diffusion coefficient value.
Therefore, it is still important to develop or optimize a
procedure to determine the both parameters simultaneously. In addition to the heat and mass transfer parameters, the analytical solutions may also be used to
determine the thermocouple locations to use in the further model validation studies. This method would be
faster and easier compared to the other methods.
Acknowledgment
The author wishes to express his gratitude to Dr. Khe
V. Chau at the University of Florida, Department of
Biological and Agricultural Engineering for his valuable
suggestions.
References
4. Conclusions
Heat (heat transfer coefficient and thermal diffusivity)
and mass (mass transfer coefficient and diffusion coefficient) transfer parameters are important to characterize,
to model and to optimize the heat and mass transfer
processes. Different approaches for experimental determination of these parameters were explained in detail.
Since these approaches with the analytical solutions require the experimental data obtained from the interested
material, they give a greater advantage over the use of
more preferred methods, e.g., the lumped system approach or use of empirical equations to determine heat
or mass transfer coefficients. As seen in these ap-
Anderson, B., & Singh, R. P. (2002). Air impingement heat transfer on
a cylindrically shaped object. The IFT Annual Meeting and Food
Expo. Presentation no; 91C-22.
Carslaw, H. S., & Jaeger, J. C. (1986). Conduction of heat in solids. New
York, NY: Oxford.
Chau, K. V. (1995–2000). Personal Communication. Department of
Biological and Agricultural Engineering, University of Florida,
Gainesville, FL.
Erdogdu, F., Balaban, M. O., & Chau, K. V. (1998). Modeling of heat
conduction in elliptical cross section: II. Adaptation to thermal
processing of shrimp. Journal of Food Engineering, 28, 241–258.
Rahman, S. (1995). Food properties handbook. Boca Raton, FL: CRC
Press Inc..
_ & Özdemir, Y. (2004). Experimental
Ünal, B., Erdoğdu, F., Ekiz, H. I.,
theory, fundamentals and mathematical evaluation of phosphate
diffusion in meats. Journal of Food Engineering, 65, 263–272.
Descargar