You are on page 1of 15

Transport in Porous Media (2005) 61: 291–305 © Springer 2005

DOI 10.1007/s11242-004-8305-8

Validation of the Local Thermal Equilibrium


Assumption in Forced Convection
of Non-Newtonian Fluids through
Porous Channels

S. A. KHASHAN1, and M. A. AL-NIMR2


1
Mechanical Engineering Department, United Arab Emirates University, PO Box 17555,
AL-AIN, U.A.E.
2
Mechanical Engineering Department, Jordan University of Science and Technology, PO
Box 3030, Irbid 22110, Jordan

(Received: 7 November 2003; accepted in final form: 15 December 2004)


Abstract. In this paper, we assess the validity of the local thermal equilibrium assumption
in the non-Newtonian forced convection flow through channels filled with porous media.
For this purpose, the problem is solved numerically using local thermal non-equilibrium
and non-Darcian models. Numerical solutions obtained over broad ranges of representa-
tive dimensionless parameters are utilized to map conditions at which the local thermal
equilibrium assumption can or cannot be employed. The circumstances of a higher mod-
ified Peclet number, a lower modified Biot number, a lower fluid-to-solid thermal con-
ductivity ratio, a lower power-law fluid index, and a lower microscopic and macroscopic
frictional flow resistance coefficients, are identified as unfavorable circumstances for the
local thermal equilibrium (LTE) condition to hold. Quantitative LTE validity maps that
reflect the proportional effect of each parameter as related to others are presented.

Key words: porous medium, porous channel, forced convection, local thermal equilib-
rium, local thermal non-equilibrium, non-Newtonian fluids, thermal behavior.

Nomenclature
asf specific interfacial area (m2 ).
b Forchheimer coefficient (m−1 ).
Bi ∗ modified Biot number, hsf asf H 2 /(εkf ).
cpf specific heat of the fluid per unit mass (J/kgK).
hsf solid-to-fluid heat transfer coefficient (W/mK).
H channel half-spacing (m).
k thermal conductivity (W/mK).
K intrinsic permeability (m2 ) of porous medium.


Author for correspondence: E-mail: skhashan@uaeu.ac.ae
292 S.A. KHASHAN AND M.A. AL-NIMR

K∗ intrinsic permeability (mn+1 ) of medium for flow of power law fluids.


kR fluid-to-solid thermal conductivity ratio, εkf /(1 − ε)ks .
n power-law index of non-Newtonian fluid.
p pressure (Pa).
P e∗ modified Peclet number, uo H /αf .
F Forchheimer dimensionless parameter, ρbK ∗ u2−n o /µ 
∗.

uo reference axial velocity based on Darcy law (m/s), n − (K ∗ /µ∗ ) (dp/dx)


u axial fluid velocity (m/s).
U dimensionless axial fluid velocity, u/uo .
T intrinsic average fluid or solid temperature (K).
x spatial axial coordinate.
y spatial transverse coordinate.
X dimensionless axial coordinate, x/H .
Y dimensionless transverse coordinate, y/H .
αf fluid thermal diffusivity, (k/ρcp )f .
ε porosity.  
δ coefficient in Equation (2),K ∗/ ε n H n+1 .
ρ fluid density (kg/m3 ).
θ dimensionless fluid or solid temperature, (T − T∞ )/(Tw − T∞ ).
Subscripts
f fluid.
s solid.
w wall.
∞ inlet.

1. Introduction
Hydrodynamic and thermal characterization of non-Newtonian fluids flow
through porous media is a topic of practical engineering importance. The
understanding of the heat transfer involved in such flows can have imme-
diate effects on many industrial applications such as those related to oil
recovery aided by thermal methods, packed bed reactors, drying or burnout
of binder systems from green compacts during colloidal processing of
ceramics, biomechanics where fluids flow through lungs and arteries, phar-
maceuticals, filtration, and fixed bed regeneration. In particular, the forced
convective heat transfer in channels filled with non-Newtonian fluid-satu-
rated porous medium represents a realistic or an idealized physical system
for many components’ cooling applications, in the aforementioned applica-
tions and in many others. The literature describing analytical and numeri-
cal studies of heat transfer for such flows has become noticeable during the
last 15 years. In his review article, Shenoy (1994) summarized the different
categories of flow and heat transfer models and the achievements in each
category. However, an extensive literature search, conducted by the authors,
revealed that all concerned studies are either restricted to Darcian flow or
VALIDATION OF THE LOCAL THERMAL EQUILIBRIUM ASSUMPTION 293

to the local thermal equilibrium (LTE) models. The Darcy flow model is
known to break down under relatively fast flow conditions. Nakayama and
Shenoy (1993) utilized a modified Brinkman–Forchheimer extension of the
Darcy model to analytically study the forced convective heat transfer for
flow between two parallel walls subjected to uniform heat flux in a highly
porous medium with a non-Newtonian power-law fluid. Their analytical
solution was attainable with the simplifying assumptions of fully developed
flow and the validity of local thermal equilibrium. Chen and Hadim (1998)
utilized the same flow model and the LTE assumption to numerically solve
the hydrodynamically and thermally developing problem without boundary
layer approximations. Alkam et al. (1998) conducted a similar numerical
study for the case of concentric annuli.
Local thermal non-equilibrium (LTNE) conditions exist due to many
obvious causes, such as the presence of distributed or concentrated heat
sources in one phase or the presence of some agency which forces differ-
ent fluid and solid boundary temperature conditions. In fact, LTNE can
be ruled out only if steady conduction, with uniform solid and fluid
thermal conductivities, is the only heat transfer process (Nield, 1998). It
also exists under other, less obvious, conditions that are mainly related
to the presence of heat convection. Many studies have examined these
conditions and the validity of the LTE assumption in general, for cases
involving Newtonian flows. One can refer, for example, to studies car-
ried out by Vafai and Sözen (1990a, b), Amiri and Vafai (1998), Nield
(1998), Al-Nimr and Kiwan (2002), Al-Nimr and Abo-Hijleh (2002), Lee
and Vafai (1998), Nield et al. (2002) and Khashan et al. (2005). The
non-Newtonian behavior, whether it is the result of the inherent char-
acter of the fluid or is due to the presence of additives, is expected
to have a significant effect on the overall flow and thermal charac-
teristics including the LTE condition. The validity of the LTE condi-
tion for the forced convection of non-Newtonian fluids in a porous
medium, and particularly that through a channel, has not been yet
considered.
Although the focus of our study is to examine the validity of the
local thermal equilibrium LTE assumption for forced convective heat
transfer of non-Newtonian fluids in a channel confined by two hor-
izontal wall planes, the findings of this study are expected to shed
light, at least qualitatively, on similar problems involving other
configurations. The departure from local thermal equilibrium is to be
captured by solving the problem numerically using a local thermal non-
equilibrium, two-medium model. A Darcy–Brinkman–Forchheimer model,
suited for highly porous media and modified for non-Newtonian flows
(Nakayama and Shenoy, 1993) is adopted to describe the hydrodynamic
behavior.
294 S.A. KHASHAN AND M.A. AL-NIMR

2. Mathematical Formulation
The analysis is carried out for a steady state, incompressible, laminar,
and two-dimensional flow in a parallel-plane walls channel filled with a
homogenous and isotropic porous medium as shown in Figure 1. The
fluid is non-Newtonian and obeys the power-law model. Flow is forced
through the channel by a constant pressure gradient and is assumed to be
hydrodynamically developed, but thermally developing. Except for the fluid
viscosity, which is dictated implicitly by the power-law model, the thermo-
physical properties of the solid and fluid phases are considered constant.
Heat generation, natural convection, thermal radiation, and viscous dissi-
pation are all neglected. We are dealing with the case in which longitudinal
(axial) conduction is negligible, so that the heat transfer process involves a
balance between longitudinal enthalpy flow (in the fluid phase), transverse
conduction and heat sources (or sinks) due to energy exchange between
fluid and solid domains. The neglecting of axial conduction is justified for
Reynolds numbers (based on the particle diameter) less than 10 (Amiri and
Vafai, 1998), a condition that is met in most practical porous flow condi-
tions. The fluid and solid phases are assumed to be in local thermal equi-
librium at the plane walls as they are uniformly subjected to the same wall
temperature.
The motion is governed by Brinkman–Forchheimer extended Darcy’s
Law. This law, as modified to power-law fluids and suited to highly porous
media with constant porosity, can be expressed as (Nakayama and Shenoy,
1993)
  
µ∗ d  du n−1 du µ∗ n dp
n   = ∗
u + ρbu2 + , (1)
ε dy dy dy K dx

Figure 1. Schematic diagram of the problem under consideration.


VALIDATION OF THE LOCAL THERMAL EQUILIBRIUM ASSUMPTION 295

where µ∗ represents the consistency of the power-law fluid. The modi-


fied permeability K ∗ , which depends on the porous structure and on the
power-law index n, and the Forchheimer constant b are determined based
on experimental correlation.
The Brinkman’s macroscopic viscous term was included in the model
because it plays an important role especially near the boundary. It rep-
resents the frictional drag forces between the fluid layers themselves. This
term is more significant in the case of viscous fluids and when there is
high velocity spatial gradient near the boundary, which is the case with
a non-Newtonian fluid flowing through a highly porous medium such
as fibrous and foam materials. For the problem under consideration, the
Brinkman term will have a significant effect in small width channels H and
when porous domains have large modified permeability K ∗ . In general, the
Brinkman term becomes more significant when the parameter marking its
relative effect δ, as defined in Equation (3), is large.
Equation (1), however, works only with constant porosity and, therefore
fails when the channeling effect due to the increase in porosity near the
wall is taking place (Nakayama and Shenoy, 1993). As for the Forchheimer
modification, which is experimentally established for Newtonian fluids, it
should be noted that this term is equally applicable to non-Newtonian flu-
ids since the microscopic inertial drag force does not depend on the viscous
stress–strain rate relationship. The use of Brinkman–Forchheimer extended
Darcy’s Law is a common practice in forced convection of Newtonian fluid
flows through porous conduits; see, for example, studies by Vafai (2002)
and Nield and Bejan (1999). An appraisal of the capabilities of Brinkman–
Forchheimer modification can be found in Nield (1991) and in Nield and
Bejan (1999).
Using the dimensionless parameters, as defined in the nomenclature, the
governing motion equation (1) is reduced to
  
d  dU n−1 dU
δ   − U n − F U 2 + 1 = 0. (2)
dY  dY  dY

The coefficients δ and F are defined, respectively, by

K∗
δ= , (3)
ε n H n+1
ρbK ∗ u2−n
o
F= . (4)
µ∗

The dimensionless macroscopic energy equations for the fluid and solid
phases, respectively, are (Nield and Bejan, 1999; Vafai, 2002):
296 S.A. KHASHAN AND M.A. AL-NIMR

∂θf 1 ∂ 2 θf Bi∗
U = ∗ + (θs − θf ) , (5)
dX Pe ∂Y 2 Pe∗
∂ 2 θs
+ Bi∗ kR (θf − θs ) = 0, (6)
∂Y 2
where the modified Péclet Pe∗ and Biot numbers Bi∗ and the fluid-to-solid
conductivity ratio kR are, respectively, defined as
uo H
Pe∗ = , (7)
αf
hsf asf H 2
Bi∗ = , (8)
εkf
εkf
kR = . (9)
(1 − ε) ks
The last term in both energy equations is due to the absence of the LTE
condition as it accounts for energy exchanges between solid and fluid
domains. The modified Biot number, as defined in this study, is related to
the equivalent Biot number Bi used by Nield et al. (2002) and by Lee and
Vafai (1998) by the relation Bi∗ = Bi(1 + kR−1 ).
The flow is assumed symmetric around the channel center, and then
only the upper half is considered for computations. The appropriate
boundary conditions are

U (X, 1) = 0, (10)

θf (X, 1) = θs (X, 1) = 1, (11)

θf (0, Y ) = θs (0, Y ) = 0, (12)


∂U ∂θf ∂θs
(X, 0) = (X, 0) = (X, 0) = 0. (13)
∂Y ∂Y ∂Y

3. Numerical Method and Solution


The steady governing equations along with their boundary condition are
solved numerically using the finite volume approach. Diffusive and convec-
tive fluxes are discretized using a second-order scheme. A stable second-
order discretization is secured for the convective term in the fluid energy
by means of the deferred correction method. In this method, fluxes discret-
ized using upwind first-order and second-order schemes are blended. The
low-order, but stable, flux contributes to the matrix coefficients, and the
difference between the two fluxes contributes to the source term so that
the second-order flux will prevail upon convergence. The source term in the
VALIDATION OF THE LOCAL THERMAL EQUILIBRIUM ASSUMPTION 297

motion equation is linearized in a way that will ensure diagonal dominance


in the corresponding discretized equation. The motion and energy fields are
uncoupled. Therefore, the motion equation is solved first iteratively until
both the change in solution for all grids and the global mass imbalance
between the inlet and the outlet of the channel fall below 10−4 . The fluid
energy followed by the solid energy equation is then solved iteratively until
the changes in the corresponding temperatures for all grids fall below 10−4 .
Systems of linearized equation are solved using the SIP method (Ferziger
and Peric, 1999).
The problem under consideration is, relatively speaking, not a compu-
tationally demanding one as it involves only three variables and as the
absence of the longitudinal diffusion in the fluid energy equation deems
this equation parabolic in the X direction, so that it is solved by march-
ing in the downstream direction. Therefore, 240 × 80 quadrilateral control
volumes are distributed uniformly along the X and Y coordinates respec-
tively. The length of the channel is equal to ten times its width. With such
fine meshing, clustering of the grids in regions near the entrance and the
wall is deemed unnecessary. Further grid refinement revealed insignificant
changes in the obtained solution.

4. Results and Discussion


The hydrodynamic and thermal parts of the aforementioned numerical
method are verified separately by comparing numerical solutions against
exact solutions available for two simplified problems. The motion problem
involving Newtonian fluid (n = 1) with negligible microscopic inertial term
(F = 0) has the exact solution of
 
cosh √Yδ
U (Y ) = 1 −  . (14)
cosh √1δ

Figure 2 shows a comparison between the numerical and exact velocity


profiles for F = 1 and n = 1. As is clear from this figure, the results are
in excellent agreement. Another verification is made by obtaining the exact
solution for the fluid energy equation under the plug flow (U = constant)
and large Bi∗ conditions. This exact solution is given as:

(−1)n  
θf (X, Y ) = 1 − 2 exp −βn2 X/(U Pe∗ ) cos βn Y , (15)
βn
n=0

where
(2n + 1) π
βn = , n = 0, 1, 2, . . .
2
298 S.A. KHASHAN AND M.A. AL-NIMR

0.001
0.8 0.01
0.1

0.6
d =1
Y
U
0.4
U (Eqn. 14)
10
n=1, F=0
0.2

0
0 0.2 0.4 0.6 0.8 1
U

Figure 2. Comparison between the numerical solution and the analytical velocity
profiles at different δ, for F = 0 and n = 1.

Also, the numerical temperature profiles are found to be in excellent


agreement with the exact ones shown in Figure 3 for different locations
along the channel length. The comparisons are made with closed form ana-
lytical solutions obtained after omitting the nonlinear terms, which mark

0.8

* *
0.6
Pe = 50 K =100 Bi =100
R

Y
0.4
X

θ f (Eqn. 15)
0.2
θf

0
0 0.2 0.4 0.6 0.8 1
qf

Figure 3. Comparison between the numerical and analytical temperature profiles at


different locations along the channel length, for a plug flow with large Bi∗ and kR .
VALIDATION OF THE LOCAL THERMAL EQUILIBRIUM ASSUMPTION 299

the non-Newtonian and microscopic inertial effects, because numerical


or experimental data for non-Newtonian fluid flow within highly porous
domains and with significant inertial term are not available in the literature.
However, it is clear from the obtained results that the numerical predictions
of the exact model approach the analytical predictions of the special case
as the two non-linearity sources diminish.
Moreover, the numerical scheme used in the study can handle the same
problem more easily if the Brinkman term is excluded.
In the following figures, the focus is on the effect of the different dimen-
sionless parameters, particularly kR , Bi∗ , Pe∗ , δ, F and n on the existence of
LTE conditions. The LTE condition is declared valid when the criterion
|θs − θf | < 0.055 everywhere in the channel. The common practice of normal-
izing the temperature departure with respect to either temperature is avoided
here as it may results in misleading values when both temperatures are effec-
tively equal or close to zero. It is worth mentioning that the thermal equi-
librium under investigation is the one based on a macroscopic scale and not
that based on the pore scale. It is believed that it is the macroscopic effects
which are expected to be of practical importance (Nield, 1998).
Figure 4 shows the effects of the modified Biot number Bi∗ , modified
Peclet number Pe∗ and fluid-to-solid conductivity ratio kR on the LTE con-
dition. Areas in the Bi∗ − Pe∗ plane mapped to the right (down) and to the
left (above) of a constant kR line can be depicted as the Bi∗ − Pe∗ condi-
tions at that kR and by which LTE or LTNE, respectively, prevails. A point
at a constant kR line refers to the minimum Biot number required at a cor-
responding Peclet number condition in order to satisfy the LTE condition.

100
n = 1.5
100

F = 0.1
LTNE d = 0.1
20
80 4
10 LTE
1
2
0.8
60
* 0.6
Pe

40 0.4

0.2
20

k =0
R

0 100 200 300 400 500


*
Bi

Figure 4. Pe∗ − Bi∗ − kR LTE validity map at n = 1.5, F = 0.1 and δ = 0.1.
300 S.A. KHASHAN AND M.A. AL-NIMR

Also, it refers to the maximum Péclet number that can be tolerated at a


corresponding Biot number in order for LTE condition to be satisfied. It
is clear that a higher kR as opposed to a lower kR , puts less constraint on
the validity of LTE, i.e. higher Péclet numbers are tolerated with lower Biot
numbers. The favorable LTE conditions of higher Biot number and higher
fluid-to-solid conductivity ratio kR are attributed to the increase in the heat
transfer, attainable at such values, between the fluid and solid phases, which
brings their temperatures closer. Higher kR implies that we have more fluid
having high thermal conductivity compared with less solid having low ther-
mal conductivity. This enables the fluid to bring the solid phase to its tem-
perature easier. Qualitative estimates for the effects of Bi∗ and kR on the
LTE conditions can be easily obtained by a scale analysis of Equations
(5) and (6). When we divide both equations by Bi∗ , it becomes clear that
as Bi∗ increases, θf approaches θs . When kR is increased, θf approaches θs
faster. It is worth mentioning that the mapping for LTE–LTNE regions
is facilitated by nesting the calculations using an utmost inner ‘Bi∗ ’ loop
located inside an outer ‘Pe∗ ’ loop which in turn is located inside an utmost
outer ‘kR ’ loop. For a selected kR value, Bi∗ and Pe∗ were incremented
sequentially in their loops by 0.5 and 0.1, respectively. However, increment-
ing Bi∗ starting from zero for all Pe∗ values would result in an unneces-
sary and exhaustive computational overhead. Instead, Bi∗ was incremented
at the new Pe∗ starting from the value at which the LTE criterion was sat-
isfied while computing on the previous Pe∗ value.
On the other hand, it is clear from the fluid energy equation that a
higher Péclet number has an unfavorable effect on the LTE condition.
As P e∗ increases, θf will feel the hot boundary condition less than the
solid porous domain feels it, so there is no way that θf = θs . Physically,
a higher Péclet number refers to the condition at which a fluid with
low thermal diffusivity is being carried through the channel at a higher
flow rate. This condition makes the fluid less capable of receiving heat
from walls and exchanging this heat to the solid phase in an efficient
manner.
Figures 5 and 6 show that the effects of higher δ and higher Forchheimer
parameter F , as expected, always extend the LTE validity conditions. As
δ and F increase, the macroscopic and microscopic frictional resistances,
respectively, increase and less fluid, which is easier to heat, is allowed to
flow through the channel. As can be seen from the same figures, the exten-
sion in the LTE condition is relatively more evident at lower kR .
Figure 7 shows the effect of the fluid index n on the LTE condition. It
is clear that the LTE validity is extended slightly further to accommodate
lower Bi∗ or higher Pe∗ values in the case of shear-thickening fluids (n > 1)
as opposed to the case of shear-thinning fluids (n < 1). However, as shown
in the same figure, the extension in LTE validity attained with higher n
VALIDATION OF THE LOCAL THERMAL EQUILIBRIUM ASSUMPTION 301
100
LTNE LTE

k =4 n = 1.5, F = 0.1
R
80
k =1
R

60

*
d = 0.1
Pe d = 1.0
40

20

k =0.2
R

0
0 100 200 300 400 500
*
Bi

Figure 5. Effect of δ on the Pe∗ − Bi∗ − kR LTE validity map at n = 1.5, F = 0.1.

100

n = 1.5, d = 0.1

80 F=10.0
k =1.0 F=0.10
k = 4.0 R
R

60

*
Pe
40

20
k =0.2
R

0
0 100 200 300 400 500
*
Bi

Figure 6. Effect of F on the Pe∗ − Bi∗ − kR LTE validity map at n = 1.5, δ = 0.1.

becomes very marginal at low Pe∗ and low Bi∗ over the total range of kR .
The favorable effects of high n on the LTE condition can be attributed to
its effect in reducing the fluid mean velocity. It is commonly known that
for shear-thickening fluids (n > 1), as the power-law index increases, the
velocity gradient near the wall decreases and the wall shear rate increases
such that the fully developed velocity profile in the channel becomes less
302 S.A. KHASHAN AND M.A. AL-NIMR

100
d = 0.1, F = 0.1
k =4.0
R
80 n=1.5
n=1.0
n=0.5

60

* k =1.0
R
Pe
max
40

20
k =0.2
R

0
0 100 200 *
300 400 500
Bi
min

Figure 7. Effect of the power-law index n on the Pe∗ − Bi∗ − kR LTE validity map
at F = 0.1, δ = 0.1.

flatter (more parabolic) with a lower mean velocity than for the Newto-
nian case. The opposite effect occurs for shear-thinning fluids (n < 1) lead-
ing to a higher mean velocity. In other words, for the same driving force
(pressure gradient), a lower flow rate of a thick (more viscous) fluid is
allowed through the porous channel compared to that of a thin fluid. This
implies that the ability of a thick fluid to attain the LTE condition is better
than that of a thin fluid. With a lower flow rate, the fluid has more time
to exchange energy with the solid domain and, therefore, to attain a tem-
perature closer to the solid temperature. This is why a thick fluid satisfies
the LTE assumption within a wider range than that of a thin fluid.
Figure 8 shows that at constant Pe∗ , the condition of satisfying a
minimum Bi ∗ for LTE to exist is by far more severe at a lower kR . This
is manifested in Figure 9 for a small kR . Over the entire ranges of power
law indices n, the effect of kR in extending the LTE validity is more dom-
inant over that of Bi∗ .
The favorable effects of high values of F , δ and n on the LTE condi-
tion can be attributed to their effect in reducing the velocity U . In gen-
eral, everything that makes U decrease will enhance the LTE condition.
One can see from Equation (5) that, as U approaches a large value, the
axial derivative ∂θf /∂x will be driven to zero, and therefore θf itself will
not change much above its specified zero inlet condition. This has an effect
similar to that of a higher Pe∗ , by which the porous solid feels wall tem-
peratures while the fluid does not.
VALIDATION OF THE LOCAL THERMAL EQUILIBRIUM ASSUMPTION 303
100

n = 0.5
80 n = 1.0
n = 2.0
*
LTE Pe = 40.0, d = 0.1, F = 0.1
60

k
R,min

40

20 LTNE

0
0 10 20 30 40
*
Bi min

Figure 8. Effect of the power-law index n on the Bi∗ − kR LTE validity map at
Pe∗ = 40, F = 0.1, δ = 0.1.

10

n = 0.5
8 n = 1.0
n = 2.0
*
6 Pe = 40.0, d = 0.1, F = 0.1
LTE
k
R,min

LTNE

0
0 100 200 300 400 500
*
Bi
min

Figure 9. Effect of the power-law index n on the Bi∗ − kR LTE validity map shown
for small kR and at Pe∗ = 40, F = 0.1, δ = 0.1.

5. Conclusion
The problem of the non-Newtonian forced convection flow in channels filled
with porous media is solved numerically using a Finite-Volume approach. A
Darcy–Brinkman–Forchheimer model, modified for non-Newtonian power-
law fluids, is adopted to describe the non-Newtonian fluid flow behavior. The
304 S.A. KHASHAN AND M.A. AL-NIMR

thermal behavior is described using a local thermal non-equilibrium LTNE,


two-medium, model.
Numerical solutions obtained over broad ranges of flow and thermal
conditions are used to clarify circumstances under which the effect of the
local thermal non-equilibrium LTNE is important and the LTE assumption
is, therefore, not valid. The favorable circumstances for the LTE assump-
tion to hold are identified. It is found that, over all ranges of power-law
fluid index n, Forchheimer parameter F and macroscopic frictional coeffi-
cient δ, the circumstances of higher Péclet numbers, lower Biot numbers,
and lower fluid-to-solid thermal conductivity ratios are all found to have
the effect of limiting the validity ranges of the LTE assumption. The effect
of a higher fluid-to-solid thermal conductivity ratio, however, is more effec-
tive in extending the LTE validity than that of a higher Biot number, i.e., at
any Péclet number condition, the requirement of satisfying a minimum Biot
number condition for LTE to exist is by far more severe at lower fluid-to-
solid thermal conductivity ratios. It is also found that the effect of higher
fluid index n is to extend the LTE validity ranges. Higher macroscopic and
microscopic frictional resistance coefficients have the effect of extending the
LTE, too. Quantitative local thermal equilibrium LTE validity maps are
presented for a wide range of hydrodynamics and thermal operating con-
ditions.

References
Abu-Hijleh, B. A., Al-Nimr, M. A. and Hader, M. A.: 2004, Thermal equilibrium in tran-
sient forced convection porous channel flow, Transport in Porous Media 57, 49–58.
Alkam, M. K., Al-Nimr, M. A. and Mousa, Z.: 1998, Forced convection of non-New-
tonian fluids in porous concentric annuli, Int. J. Numerical Methods Heat Fluid Flow
8(6), 703–716.
Al-Nimr, M. A. and Abo-Hijleh, B. A.: 2002, Validation of thermal equilibrium assump-
tion in transient forced convection flow in porous channel, Transport in Porous Media
49, 127–138.
Al-Nimr, M. A. and Kiwan, S.: 2002, Examination of thermal equilibrium assumption in
periodic forced convection in a porous channel, J. Porous Media 5(1), 35–40.
Amiri, A. and Vafai, K.: 1994, Analysis of dispersion effects and non-thermal equilibrium
non-Darcian, variable porosity incompressible flow through porous media, Int. J. Heat
Mass Transfer 37, 939–954.
Amiri, A. and Vafai, K.: 1998, Transient analysis of incompressible flow through a packed
bed, Int. J. Heat Mass Transfer 41, 4259–4279.
Chen, G. and Hadim, H. A.: 1998, Forced convection of a power-law fluid in a porous
channel – numerical solutions, Heat Mass Transfer 34, 321–228.
Ferziger, J. H. and Peric, M.: 1999, Computational Methods for Fluid Dynamics, Springer,
Berlin, Heidelberg, New York.
Khashan, S. A., Amiri, A. and Al-Nimr, M. A.: 2005, Assessment of the Local
Thermal Non-Equilibrium Condition in Developing Forced Convection Flows through
Fluid-Saturated Porous Tubes, Applied Thermal Engineering, in press.
VALIDATION OF THE LOCAL THERMAL EQUILIBRIUM ASSUMPTION 305

Lee, D. Y. and Vafai, K.: 1998, Analytical characterization and conceptual assessment of
solid and fluid temperature differentials in porous media, Int. J. Heat Mass Transfer
42, 423–435.
Nakayama, A. and Shenoy, A. V.: 1993, Non-Darcy forced convective heat transfer in a
channel embedded in a non-Newtonian inelastic fluid-saturated porous medium, Cana-
dian J. Chem. Eng. 71, 168–173.
Nield, D. A.: 1991, The limitations of the Brinkman-Forchheimer equation in modeling
flow in a saturated porous medium and at an interface. Int. J. Heat Fluid Flow 12,
269–272.
Nield, D. A.: 1998, Effect of local thermal non-equilibrium in steady convective processes
in saturated porous medium: Forced convection in a channel. J. Porous Media 1, 181–
186.
Nield, D. A. and Bejan, A.: 1999, Convection in Porous Media, Springer, New York.
Nield, D. A., Kuznetsov, A. V. and Xiong Ming: 2002, Effect of local thermal no-equi-
librium on thermally developing forced convection in a porous medium, Int. J. Heat
Mass Transfer 45, 4949–4955.
Shenoy, A. V.: 1994, Non-Newtonian fluid heat transfer in porous media, Adv. Heat
Transfer 24, 101–190.
Vafai K. (ed.): 2002, Handbook of Porous Media, Marcel Dekker, New York.
Vafai, K. and Sözen, M.: 1990a, Analysis of energy and momentum transport for fluid
flow through a porous bed, ASME J. Heat Transfer 112, 690–699.
Vafai, K. and Sözen, M.: 1990b, An investigation of a latent heat storage porous bed
and condensing flow through it, ASME J. Heat Transfer 112, 1014–1022.

You might also like