Sie sind auf Seite 1von 15

This article was downloaded by: [University of Waterloo]

On: 11 October 2014, At: 21:53


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Numerical Heat Transfer, Part A: Applications: An


International Journal of Computation and Methodology
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/unht20

EFFECTS OF BOUNDARY CONDITIONS ON NON-DARCIAN


HEAT TRANSFER THROUGH POROUS MEDIA AND
EXPERIMENTAL COMPARISONS
a

A. Amiri , K. Vafai & T. M. Kuzay

Department of Mechanical Engineering , The Ohio State University , Columbus, Ohio,


43210, USA
b

Argonne National Laboratory , Argonne, Illinois, 60439, USA


Published online: 29 Oct 2007.

To cite this article: A. Amiri , K. Vafai & T. M. Kuzay (1995) EFFECTS OF BOUNDARY CONDITIONS ON NON-DARCIAN HEAT
TRANSFER THROUGH POROUS MEDIA AND EXPERIMENTAL COMPARISONS, Numerical Heat Transfer, Part A: Applications: An
International Journal of Computation and Methodology, 27:6, 651-664, DOI: 10.1080/10407789508913724
To link to this article: http://dx.doi.org/10.1080/10407789508913724

PLEASE SCROLL DOWN FOR ARTICLE


Taylor & Francis makes every effort to ensure the accuracy of all the information (the Content) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.
This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions

EFFECTS OF BOUNDARY CONDITIONS O N


NON-DARCIAN HEAT TRANSFER THROUGH
POROUS MEDIA A N D EXPERIMENTAL
COMPARISONS
A. Amiri and K. Vafai
Downloaded by [University of Waterloo] at 21:53 11 October 2014

Department of Mechanical Engineering, The Ohio State Uniuersity,


Columbus, Ohio 43210, USA

T. M. Kumy
Argonne National Laboratory, Argonne, lllinois 60439, USA
The present work centers around the numerical simuhtion offorced conuectioe incompressible flow thmugh porous beds. Inertial as weU as uiscous effects are considered i n the
momenhrm equation. The mathemafr'cal model for energy transport was based on the
two-phase equation model, which does not employ local thermal equilibrium assumption
between the Jluid and the solid phases. The transport processes for hvo diflerent types of
boundnry conditions are studied. 7'he analysis was pegarmed i n terms of nondimensional
parameters that successfuUy cast together OU the pertinent influencing effects. Comparisons
were made between our numerical findings and experimertlal resub. OueraU, the
comparisons that were madefor the constant waU h e d m boundary condition display good
agreement.

INTRODUCTION

Forced convection heat transfer in porous media has been extensively investigated. The phenomenal enhancement in heat transfer level produced by using
porous materials as transport media has stimulated significant interest for researchers to develop rigorous models to simulate thermal behavior in porous
media. Typical examples of applications that utilize porous media include absorption and adsorption processes, packed filters, pebble-type heat exchangers, and
energy storage units. In addition, utilization of porous insertions in high heat flux
applications as a technique to augment heat transfer has proved to be a promising
research field, as evidenced by the experimental investigations of metallic porous
media [I, 21. These experimental results have shown phenomenal increases in the
level of heat transfer.
There exists a significant amount of published literature concerned, in the
main, with the problem of forced convective flow through bounded porous media.
Received 28 April 1994; accepted 10 August 1994.
Address correspondence to Dr. Kambiz Vafai, Department of Mechanical Engineering, The
Ohio State University. Columbus, OH 43210-1107, USA.

Numerical Heat Transfer, Part A, 27:651-664,1995


Copyright O 1995 Taylor & Francis
1040-7782/95 $10.00 + .OO

A. AMlRI ET AL.

652

NOMENCLATURE

"

.t

C~

d~
dp/&
Da

Dh

Downloaded by [University of Waterloo] at 21:53 11 October 2014

F
hSf
H
i

I
J
k
K

L
LTE

Nu
P
Pr
9,
Re
Re,
I

T
U

specific surface area of the packed


bed, mspecific heat at constant
pressure, J kg- Kpore diameter, m
pressure gradient, N m-3
Darcy number (K/H2)
hydraulic diameter of the
channel, m
geometric function
fluid-to-solid heat transfer
coefficient, W m - 2 K-I
height of the packed bed, m
index for x coordinate
index for y coordinate
unit vector oriented along the
pore velocity vector
thermal conductivity, W m-' K-'
permeability, m2
length of the packed bed, m
local thermal equilibrium
Nusselt number
pressure
Prandtl number (= p r c p r / k , )
heat flux, W m - 2
Reynolds number ( = p , u , D h / p f )
particle Reynolds
number ( = p f u c d p / p f )
time, s
temperature, K
velocity component in the
x direction, m s''

'

'

'

u
x, Y
a

A,
M
P

velocity vector, m s-'


Cartesian coordinates, m
thermal diffusivity, m2 s - '
porosity
inertia parameter

I=

~~r"~F(~,H/pr)l

kinematic viscosity, kgm-I s-I


density, kgm-'

Subscripts
d
ed
eff
f
m
s
w
x

Y
0

dispersion
inlet
effective property
fluid
mean
solid
wall
x component
y component
reference

Superscripts
f
s

fluid
solid

Symbol

"local volume average"


of a quantity

The complicated and interesting phenomena associated with porous beds provide a
wide area for fruitful discussion. In the work of Vafai and Tien [3], the inertial and
impermeable boundary effects on momentum and energy transport were discussed
in great detail. The investigation provided a simple characterization scheme for
interpreting the applicability of the familiar Darcy's law for various flow conditions
and bed configurations.
The fluid-to-solid heat transfer coefficient has received considerable attention in terms of experimental investigations to formulate its quantity for porous
beds of different geometries and configurations. Studies by Barker [4] and Wakao
et al. [5,6] can be regarded as examples of investigations to formulate the
fluid-to-solid heat transfer coefficient. Moreover, the flow through complicated
structures such as porous beds introduces flow disturbance, mixing, and recirculation of local fluid streams as the fluid passes through the bed. This phenomenon is

Downloaded by [University of Waterloo] at 21:53 11 October 2014

NON-DARCIAN HEAT TRANSFER THROUGH POROUS MEDIA

653

referred to in the literature as thermal dispersion. Extensive studies have been


conducted to establish empirical correlations to formulate the effective thermal
conductivities in porous beds in general and packed beds in particular. Investigations by Yagi and Kunii [7,8], Yagi et al. [9], and Wakao et al. [5,6] reported
experiments on measurement of their quantities.
Most of the research related to forced convective flow in porous media is
constrained by the assumption of the presence of local thermal equilibrium (LTE)
between the fluid and the solid phases. Studies such as those of Cheng and
Vortmeyer [lo], Hunt and Tien [Ill, and Poulikakos and Renken [I21 have
employed such an assumption. A rigorous model has been developed by Vafai and
Sozen [13] for forced convective compressible flow. The proposed model takes into
consideration the local temperature variation between the fluid and solid phases in
an attempt to attain a better understanding of the thermal characteristics of porous
beds. In a recent paper, the authors have reported a study on steady state,
incompressible flow through a packed bed [14]. In that work, the non-Darcian
effects, the variable porosity effects, and the dispersion effects were analyzed.
Furthermore, the LTE assumption and the one-dimensional approach were discussed and expressed in terms of some governing parameters and were presented
in an integrated form for different flow conditions and porous media configurations.
The present investigation is concerned with the steady state analysis of
incompressible flow through porous beds. Boundaly conditions of constant wall
temperature and constant wall heat flux are employed, and their effects on the
thermal response of the bed are discussed. In addition, the numerical results for
constant wall heat flux are compared with the experimental findings of Kuzay et al.
[2]. This comparison is essential, since the available literature, to the authors'
knowledge, has not been able to simulate proposed models to the latest experimental results.

ANALYSIS

The analysis is made for an incompressible, two-dimensional channel filled


with a metallic porous material of constant porosity. Water is used as the heat
transfer medium. The schematic diagram of the problem is shown in Figure 1. The
length and height of the bed were chosen to be 550 mm and 9.5 mm, respectively.
These physical values are representative dimensions of the experimental investigation [21.

Figure 1. Schematic diagram of the prob-

A. AMIN El'AL.

Governing Equations

The system of governing equations is presented in the following vectorial


form [3,13]:
Continuity

V(u)

Downloaded by [University of Waterloo] at 21:53 11 October 2014

Momentum

Fluid phase energy

Solid phase energy

where u is the velocity vector, p, the fluid density, the average porosity, pf the
dynamic viscosity of the fluid, K the permeability of the porous bed, F the
geometric function, J = u,/lupl the unit vector oriented along the pore velocity
vector up,( P)' the average pressure read off a pressure gage, c,, the fluid-specific
heat, (T,)' the intrinsic phase average of the fluid temperature k,, the fluid
effective thermal conductivity, h , the fluid-to-solid heat transfer coefficient, a,,
the specific surface area of the porous bed, p, the solid density, c,, the solid
specific heat, (T,)" the intrinsic phase average of the solid temperature, and k,,,,
the solid effective thermal conductivity.
The specific surface area of the bed is formulated as [13,15]

where d p is the pore diameter. The fluid-to-solid heat transfer coefficient was

NON-DARCIAN HEAT TRANSFER THROUGH POROUS MEDIA

655

based on an empirical correlation established by Wakao et al. [5] and is expressed


in the following form:

Downloaded by [University of Waterloo] at 21:53 11 October 2014

where Pr is Prandtl number and u is the velocity component in the x direction.


The thermal dispersion effect was treated as a diffusive component added to the
fluid longitudinal and lateral effective thermal conductivities [Ill. The empirical
correlation developed by Wakao and Kaguei [6] is employed in this study to model
the fluid-effective conductivities:

where k, and k , are the fluid and solid thermal conductivities, respectively. It is
worth noting that Pr is assumed constant due to the small temperature range
considered in this study.
The local heat flux is cast in a dimensionless form, namely, the Nusselt
number. The definition of Nu is based on the hydraulic diameter:

In the above equation, q, is the wall heat flux, Dh is the hydraulic diameter, Tw is
the wall temperature, and T,, is the mixed mean fluid temperature defined in a
manner similar to that for classical channel flows. It is important to point out that
the total heat flux generated for the case of a constant wall temperature is
computed by summing the heat flux generated from both the fluid phase and the
solid phase.
The particle Reynolds number Re, and the Darcy number Da were utilized
to discuss the energy transport in porous beds. In addition, different permeability
values and different materials were considered for the solid phase. The numerical
computations were carried out for Da =
and 5 x
The main criterion in
selecting the solid phase materials was to obtain a wide range of values for the
solid-to-fluid thermal diffusivity ratio a,/af since the thermophysical properties of
the porous material may differ widely from one application to another. Table 1
presents the thermophysical properties of the materials considered in the present
investigation.
For completeness of the mathematical model, one needs to specify the
porosity of the bed and the pore diameter. For practical reasons, the latter is taken
to be 1 mm, whereas the former is taken to be 0.95 and the geometric function is

A. AMIRl ET AL.

656

Table 1. Thermophysical properties of the considered materials

Downloaded by [University of Waterloo] at 21:53 11 October 2014

Water
Solid phase
Copper
Bronze
Sandstone

Density,
kg m - 3

Specific heat,
J ~ ~ K- - II

997

4179

8933
8800
2200

385
420
710

Thermal
conductivity,

w m-l

K-I

0.613

Dynamic
viscosity,
x104 kgm-ls-l
8.55

401
52
1.83

taken as 0.08. On the other hand, the permeability and the geometric function used
for comparison with the experimental results were obtained directly from the
experimental investigations of Kuzay et al. [2] in a manner similar to that reported
by Vafai and Tien [16].
Boundary Conditions

In the problem under investigation, a no-slip boundary condition is imposed


at the solid boundary. Furthermore, the fluid and the solid phases are at the same
uniform temperature at the inlet of the bed. That is (dropping the local volume
average sign for convenience);

where T, is the entrance temperature. Both constant wall temperature and


constant wall heat flux were analyzed in the present analysis. The boundary
conditions may be expressed in the following form:
Constant wall temperature at the top and bottom walls

Constant wall heat flux at the top and bottom boundaries

for both Tfand T,.


The choice of the boundary condition for the wall heat flux case using the
two-equation energy model is not a trivial task. To our knowledge, no discussion
on the proper presentation of the wall heat flux boundary condition for the two-

Downloaded by [University of Waterloo] at 21:53 11 October 2014

NON-DARCIAN HEAT TRANSFER THROUGH POROUS MEDIA

657

equation energy mode has been reported. The wall heat flux boundary condition
may be viewed in two different ways. The first is to assume that each representative
elementary volume (which contains both fluid and solid phases) at the wall surface
receives a prescribed heat flux that is equal to the wall heat flux q,. As a result,
the q, will be divided between the two phases on the basis of the physical values of
their effective conductivities and their corresponding temperature gradients. The
second approach is to assume that each of the individual phases at the wall surface
will receive an equal amount of heat flux q,. Due to the non-slip boundary
condition at the wall, the convective mode between the two phases tends to vanish
at the wall surface. Consequently, the LTE assumption prevails close the wall, and
as a result, it is reasonable to assume that each of the individual phases at the wall
surface will receive an equal amount of heat flux q,. The numerical experimentation in this study was found to validate this argument, and it was therefore
implemented in this work.
SOLUTION METHODOLOGY
Equations (1)-(4) were solved by using the finite difference method. The
momentum and the energy equations were solved separately, since they are not
coupled. Central differencing was used for the spatial derivatives except for the
convective term in Eq. (31, for which the upwind differencing scheme was employed. The momentum equation was first solved by a tridiagonal matrix algorithm
after linearizing the nonlinear term. Next, the energy equations were solved after
specifying the exit boundary condition. A three-point backward differencing was
employed for the spatial x derivatives at the grid points on the right boundary. The
strong parabolic nature of the forced convective flows validates this procedure.
Furthermore, this procedure was validated by a systematic increase in the computational length of the bed, until the numerical results within the physical domain
were no longer affected by further increase in the computational length. The
steady state solutions of these equations were then obtained. The source terms that
appear in Eqs. (3) and (4) were updated after each iteration due to their temperature dependency.
Variable grid size was implemented in the y direction, while constant grid
size was used in the x direction. A fine, equally spaced grid size was positioned
within 16% of the total height from each solid boundary, while a relatively coarser
equally spaced grid size was used for the core region. It was ensured that the
results were grid independent. This was done through a systematic decrease in grid
size until a grid-independent result was obtained. A 101 x 101 grid configuration
was found to satisfy this criterion. The results were assumed to have converged
when the temperature values for the fluid and the solid phases for two consecutive
iterations were below the convergence criterion of 10-lo.
RESULTS AND DISCUSSION
In order to examine the accuracy of our numerical results, the numerical
results were compared against the most closely related analytical solutions. Limiting cases of analytical solutions were used for this purpose. The necessary adjust-

A. AMlRI

658

ET AL.

Downloaded by [University of Waterloo] at 21:53 11 October 2014

ments were made to our numerical code to reduce it to a system equivalent to the
simplified available solutions. O u r numerical finding for the velocity field was first
compared with the analytical solution of Vafai and Kim [17]. The comparison is
depicted in Figure 2 a . Next, the temperature distribution was compared with the
analytical results of Vafai and Kim [17] for the case of constant wall heat flux, and
with the analytical solution of Vafai and Thiyagaraja 1181 for the case of constant
wall temperature. Figures 2b and 2c demonstrate such comparisons. It is worth
noting that all of the above comparisons were made in terms of the dimensionless

ra5
04

03

- rrw-

02
01

--m

aW.lo

02

04

"tu.

oe

aa

Figure 2 0 . Cornpanson of velocity profile of the present work w ~ t hthe analytical solution of Vafa~and Kim 1171.

Figure Zb. Comparison of the fully developed temperature profile of the present
work with the analytical solution of Vafai
and Kim 1171 for q, = const.

Figure Zc. Comparison of the fully developed temperature profile of the present
work the analytical solution of Vafai and
Thiyagaraja 1181 for T, = const.

Downloaded by [University of Waterloo] at 21:53 11 October 2014

NON-DARCIAN HEAT TRANSFER THROUGH POROUS MEDIA

659

variables that appear in the above-mentioned studies. As may be seen from Figure
2, all comparisons display excellent agreement.
The results of this study focus on the thermal response of the porous bed to a
number of influencing parameters such as the particle Reynolds number, the Darcy
number, and the solid-to-fluid thermal diffusivity ratio. The results are presented
in terms of Nusselt number for boundary conditions of constant wall temperature
and constant wall heat flux. However, since the experimental results are based on a
constant wall heat flux, the constant wall heat flux boundary condition is considered for the experimental comparisons. In addition, the wall temperature and the
average friction factor will also be highlighted when comparisons are made with
experimental results. The average friction factor is defined in a manner similar to
that for classical channel flows:

In the above equation, u , is the mixed mean fluid velocity defined in a manner
similar to that for classical channel flows.
Effect of t h e Particle Reynolds Number
In order to determine the effect of the particle Reynolds number Re, on the
Nusselt number Nu, cases with different Darcy number Da were compared. These
comparisons were performed using bronze as the solid phase. Figures 3a and 36
show the effect of the Re, on Nu for the two different boundary conditions under
consideration. As expected, it is observed that as Re, increases, the thermal
boundary layer becomes thinner, resulting in more energy being carried away by
convection than by conduction, and as a result, larger values of Nu are encountered. The Nu values are found to be much larger for the constant wall temperature than for the constant wall heat flux. This is primarily due to the way that the
Nu is calculated from the two-equation model. The contribution of the solid phase
to the total heat flux generated, which is incorporated in the Nu calculation for the
constant wall temperature, has significantly increased the magnitude of Nu.
Effect of t h e Darcy Number
The effect of Da was examined in a manner similar to that carried out in
determining the effect of Re,. Figure 4 depicts the effect of Da for the case of
bronze as the solid phase. The thickness of the momentum boundary layer for a
low-permeability porous medium is a hnction of Da only, while it becomes a
function of Da and the inertia parameter A , as the permeability increases [17]. In
addition, the flow inside the channel becomes more of a slug flow as Da decreases,
causing the magnitude of Nu to increase. Thus, the magnitude of Nu reaches its
asymptotic maximum value as Da decreases. It is important to point out that the
maximum asymptotic value, as illustrated in Figure 4, is different for different Re,.

Downloaded by [University of Waterloo] at 21:53 11 October 2014

v-.mm=-

?rn

(b)

,I

Figure 3. Variation of the Nusselt number for bronze with Da = lo-' and 5 X
lo-' for (a) constant wall temperature
and ( b ) constant wall heat flux.

The differences between the asymptotic maximum values are primarily due to the
dependence of the dispersion parameter on the velocity vector, which as a result,
affect the magnitude of the thermal resistance by a considerable margin, depending on the magnitude of the velocity.

Figure 4. Variation of the Nusselt number for bronze with Rep = 40, 100, and
200.

Downloaded by [University of Waterloo] at 21:53 11 October 2014

NON-DARCIAN HEAT TRANSFER THROUGH POROUS MEDIA

Figure 5. Variation of the Nusselt number for copper, bronze, and sandstone for
Da = 5 x

Effect of t h e Solid-to-Fluid Thermal Diffusivity Ratio


Figure 5 illustrates the effects of the solid-to-fluid thermal diffusivity ratio for
a f i e d Da equal to 5 X lo-'. Nu is found to increase as this ratio increases. It is
important to recapitulate that the magnitude of Nu is higher for the constant wall
temperature than for the constant wall heat flux, since the heat flux computed for
the constant wall temperrtture is the summation of the heat flux generated by the
fluid phase and the solid phase, as explained above.
Experimental Comparisons
The final objective of this work is to compare our numerical findings with the
experimental results reported by Kuzay et al. [2]. The experiment was conducted
using a square channel with constant heat flux imposed on the walls. Different
copper mesh configurations were used to investigate their influence on the channel's thermal response. The comparisons in this study will be made with the formed
mesh and the foam mesh type fillers, as they constitute regular porous media. It is
worth noting that the comparisons were made with supplementary detailed results
provided by the authors [2]. The numerical investigation was performed after
specifying some essential quantities. The geometric functions and the associated
permeabilities for the two examined meshes were obtained from the data of the
flow measurements at various pressure differences across the test section. The
above-mentioned quantities were estimated from the following equation:

Table 2 lists the numerical values of the geometric functions and the corresponding
permeabilities (presented here in terms of Da). For the sake of quantitative
comparisons, the hydraulic diameter was set equal to the diameter of the square
channel.

A. AMIRI ET AL.
Table 2. Characteristics for the different mesh types

Mesh

Downloaded by [University of Waterloo] at 21:53 11 October 2014

Formed
Foam

Porosity, %

Darcy number

Geometric
function

85
95

2.2 x l o 4
5.1 X lo-'

0.01 1
0.010

Figure 6. Comparison of the friction factor predictions of the present work with
the experimental data reported by Kuzay
et al. [21.

The friction factor is presented first, as depicted in Figure 6. Next, the


variation of the wall temperature at Re = 1.5 X lo4 is shown in Figure 7. As may
be seen from Figure 7, the experimental data are in good agreement with the
numerical findings. The comparisons are completed by demonstrating the variation
of the reduced Nusselt number, N u / P ~ ' . ~versus
,
the Reynolds number as shown in
Figure 8. The maximum relative difference between the experimental data and our
numerical results occurs at about Re = 4300. However, it can be seen from Figure
8 that the overall view of the results shows good agreement. Although the choice of
the volume element for the averaging process applies only to two-dimensional
problems in a strict sense [3], the outcomes of the present model as shown do not
indicate that the two-dimensional-mode1 results can be extended to compare
favorably with the presented experimental results.

Figure 7. Comparison of the wall temperature predictions of the present work


with the experimental data reported by
Kuzay et al. 121 at Re = 1.5 X lo4.

NON-DARCIAN HEAT TRANSFER THROUGH POROUS MEDIA

Downloaded by [University of Waterloo] at 21:53 11 October 2014

Figure 8. Comparison of the reduced

Nusselt number predictions of the present work with the experimental data
reported by Kuzay et al. [2].

CONCLUSIONS
T h e problem of forced convection in a channel filled with a porous material
is analyzed. T h e two-equation model is utilized to perform the analysis. Both the
constant wall temperature and the constant wall heat flux were analyzed and it is
shown that the constant wall temperature has a n edge. T h e analysis was performed
in terms of three nondimensional parameters that successfully cast together all the
pertinent influencing effects. T h e investigation was completed by comparing the
numerical analysis with a recent experimental study. T h e comparisons display good
agreement for the Nusselt number predictions and the wall temperature variation.

REFERENCES
1. F. E. Megerlin, R. W. Murphy, and A. E. Bergles, Augmentation of Heat Transfer in
Tubes by Use of Mesh and Brush Inserts, ASME J. Hear Transfer, pp. 145-151, 1974.
2. T. M. Kuzay, J. T. Collins, A. M. Khounsary, and G. Morales, Enhanced Heat Transfer
with Metal-Wool-Filled Tubes, Proc. ASME / JSME Thermal Eng. Cont, pp. 145-151,
1991.
3. K. Vafai and C. L. Tien, Boundary and Inertia Effects on Flow and Heat Transfer in
Porous Media, Int. J. Heat Mass Transfer, vol. 108, pp. 195-203, 1981.
4. J. J. Barker, Heat Transfer in Packed Beds, Ind. Eng. Chem., vol. 57, no. 4, pp. 43-51,
1965.
5. N. Wakao, S. Kaguei, and T. Funazkri, Effect of Fluid Dispersion Coefficients on
Particle-to-Fluid Heat Transfer Coefficients in Packed Beds, Chem. Eng. Sci., vol. 34, pp.
325-336, 1979.
6. N. Wakao and S. Kaguei, Heat and Mass Transfer in Packed Beds, Gordon and Breach,
New York, 1982.
7. S. Yagi and D. Kunii, Studies on Effective Thermal Conductivities in Packed Beds,
AIChE J., vol. 3, no. 3, pp. 373-381, 1957.
8. S. Yagi and D. Kunii, Studies on Heat Transfer near Wall Surface in Packed Beds,
AIChE J., vol. 61, no. 1, pp. 97-104, 1960.

Downloaded by [University of Waterloo] at 21:53 11 October 2014

664

A. AMlRl ET AL.

9. S. Yagi, D. Kunii, and N. Wakao, Studies on Axial Effective Conductivities in Packed


Beds, AIChE J., vol. 6, no. 4, pp. 543-546, 1960.
10. P. Cheng and D. Vortmeyer, Transverse Thermal Dispersion and Wall Channelling in a
Packed Bed with Forced Convective Flow, Chem. Eng. Sci, vol. 43, no. 9, pp. 2523-2532,
1988.
1 1 . M. L. Hunt and C. L. Tien, Non-Darcian Convection in Cylindrical Packed Beds, J. Heat
Transfer, vol. 110, pp. 378-384, 1988.
12. D. Poulikakos and K. J. Renken, Forced Convection in a Channel Filled with Porous
Medium, Including the Effects of Flow Inertia, Variable Porosity, and Brinkman
Friction, Int. J. Heat Mass Transfer, vol. 109, pp. 880-888, 1987.
13. K. Vafai and M. Sozen, Analysis of Energy and Momentum Transport for Fluid Flow
through a Porous Bed, J. Heat Transfer, vol. 112, pp. 690-699, 1990.
14. A. Amiri and K. Vafai, Analysis of Dispersion Effects and Non-Thermal Equilibrium,
Non-Darcian, Variable Porosity Incompressible Flow through Porous Media, Inr. J. Hear
Mass Transfer, vol. 37, no. 6, pp. 939-954, 1994.
15. F. A. L. Dullien, Porom Media Fluid TmnspoH andPore Structure, Academic, San Diego,
Calif., 1979.
16. K. Vafai and C. L.Tien, Boundary and Inertia Effects on Convective Mass Transfer in
Porous Media, Int. J. Heat Mass Transfer, vol. 25, no. 8, pp. 1183-1190, 1982.
17. K. Vafai and S. J. Kim, Forced Convection in a Channel Filled with a Porous Medium:
An Exact Solution, J. Heat Transfer, vol. 111, pp. 1103-1106, 1989.
18. K. Vafai and R. Thiyagarja, Analysis of Flow and Heat Transfer at the Interface Region
of a Porous Medium, Int. J. Heat Mass Transfer, vol. 30, no. 7, pp. 1391-1405, 1987.

Das könnte Ihnen auch gefallen