You are on page 1of 11

Applied Thermal Engineering 137 (2018) 101–111

Contents lists available at ScienceDirect

Applied Thermal Engineering


journal homepage: www.elsevier.com/locate/apthermeng

Research Paper

Forced convective heat transfer of tubes sintered with partially-filled T


gradient metal foams (GMFs) considering local thermal non-equilibrium
effect

Z.G. Xu , J. Qin, X. Zhou, H.J. Xu
School of Mechanical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China

H I GH L IG H T S

• Gradient metal foams’ forced convective heat transfer is numerically studied.


• Morphology gradients deeply influence solid and fluid temperature profile.
• The velocity decreases with increasing interfacial radius.
• Porosity gradient effect on friction factor increases with increasing pore-density.

A R T I C LE I N FO A B S T R A C T

Keywords: Gradient metal foams (GMFs) are novel porous structures and have great potential in heat transfer performance.
Gradient metal foam In the present study, fully developed forced convective heat transfer in tubes sintered with partially-filled GMFs
Forced convective heat transfer is numerically investigated. The inner surface of the tubes is subjected to constant heat flux. In the GMF region,
Local thermal non-equilibrium the Brinkman extended Darcy flow model and the local thermal non-equilibrium (LTNE) model are used to
Flow resistance
predict fluid and thermal transport. At the layer-layer interface in GMF, fluid temperature and solid energy are
assumed to be continuous. At the GMF-fluid interface, no-slip coupling conditions are used to couple flow and
heat transfer of the foam and free regions. Velocity distribution, temperature profile, the friction factor and
Nusselt number in partially-filled GMF tubes are calculated. The results show that the heat transfer performance
and flow resistance of GMFs are heavily dependent on porosity, pore density and GMF thickness gradients.

1. Introduction results improving boiling heat transfer are unsatisfactory. Moreover,


these methods destroy UMF original structures. To fully make use of the
Compared with traditional heat exchangers, metal foam is a special heat transfer potential of metal foams, the present authors developed
man-made porous structure providing larger surface area, higher solid the novel porous structures called gradient metal foams (GFMs) [3,4],
thermal conductivity and stronger flow-mixing capability, and can en- which not only provide reasonable space for the expanding liquids or
hance heat transfer considerably, which can be made into compact heat escaping growing bubbles due to enlarging pores vertical to a heating
sinks et al. Metal foam’s heat transfer performance is affected by not surface, but also have the advantages of high surface area, disturbing
only morphology parameters (porosity, pore density, thickness, etc.), liquid ability and reducing the thickness of boundary layer. The re-
but also working liquid properties. However, under boiling condition, if search results [3,4] showed that, GFMs significantly enhance pool
open-celled uniform metal foams (UMFs) are thick or dense, the re- boiling heat transfer compared to single-layer uniform foams and the
sistance slowing down bubble escaping velocity is particularly obvious, enhancement is heavily dependent on GFM layer thickness and pore
which leads to boiling deterioration, and even worse, boiling heat density gradient. When the heat flux is less than a certain value, pool
transfer performances of some UMFs are lower than the smooth surface. boiling heat transfer performance of copper-nickel foam is better than
To improve boiling heat transfer performance of UMFs, the present nickel-nickel foam because of the former’s higher thermal conductivity.
authors [1,2] added surfactant into working liquid or cut grooves in The enhancement degree of the three-layer GMF is lower than the
uniform metal foams to facilitate the boiling process. However, the double-layer GMF due to serious bubble escaping resistance. As is


Corresponding author.
E-mail address: zhiguoxu@sjtu.edu.cn (Z.G. Xu).

https://doi.org/10.1016/j.applthermaleng.2018.03.074
Received 8 February 2017; Received in revised form 22 July 2017; Accepted 21 March 2018
Available online 27 March 2018
1359-4311/ © 2018 Elsevier Ltd. All rights reserved.
Z.G. Xu et al. Applied Thermal Engineering 137 (2018) 101–111

Nomenclature Ro dimensionless tube radius


T temperature, K
asf specific surface area, m2·m−3 u velocity, m·s−1
c specific heat capacity, J·kg−1·K−1 U dimensionless velocity
df fiber diameter, m UMF uniform metal foam
dp pore diameter, m x position along x direction, m
Da Darcy number
f friction factor Greek symbols
GMF gradient metal foam
h heat transfer coefficient, W·m−2·K−1 ε porosity
hsf local heat transfer coefficient at solid-fluid contacting ω pore density, PPI
surface, W·m−2·K−1 μ dynamic viscosity, kg·m−1·s−1
k thermal conductivity, W·m−1·K−1 θ dimensionless temperature
ke effective thermal conductivity, W·m−1·K−1 ρ density, kg·m−3
K permeability, m2 φ polar angle, rad
Nu Nusselt number
p pressure, N·m−2 Subscripts
P dimensionless pressure drop
Pr Prandtl number b bulk
qw heat flux exposed to the tube wall, W·m−2 e effective/equivalent
r radius, m f fluid/fiber/foam
ri interfacial radius for the foam-fluid interface, m g gradient
rm interfacial radius for the interface between two form i interfacial
layers of the gradient metal foam, m m mean
ro tube radius, m s solid
Re Reynolds number w wall
Red local Reynolds number for flow around cylinder x x position
Re K permeability Reynolds number
R dimensionless radius Others
Ri dimensionless interfacial radius
Rm dimensionless interfacial radius for the interface between 〈〉 volume-averaging value
two form layers of the gradient metal foam

known to us all, in the forced convective fluid flow systems, heat are used to characterize the thermo-flow fields inside the porous re-
transfer can be enhanced by porous structures because of strong mixing gions. The results show that an increase in the solid-fluid interfacial
capacity which generates turbulent flow and different stream patterns. heat exchange results in a decrease in the temperatures difference be-
GMFs are the novel man-made porous media and their forced heat tween the solid and fluid phases for fixed Reynolds number. The present
transfer performance should be fully investigated. author Xu [22] numerically investigated forced convective heat transfer
Carrying out experiments and conducting numerical study are the of nanofluid in UMF duct. Forchheimer model considering the quad-
two key methods for investigating convective heat transfer [5,6]. Fluid ratic term is used for the momentum equation and local thermal non-
flow and heat transfer performance for UMFs has been extensively equilibrium (LTNE) model is used for the energy equations. The results
studied by many researchers [7–15]. Hsieh et al. [16] found that Nus- showed that pressure drop and Nusselt number for nanofluid in UMFs
selt number of uniform aluminum-foam in a pipe increases with the increase with increasing nanoparticle volume fraction. To numerically
increase pore density. Nazari et al. [17] experimentally investigated the investigate flow and heat transfer performance in metal foams, the re-
forced convective heat transfer due to flow of Al2O3/Water nanofluid searchers adopted local thermal non-equilibrium (LTNE) model and
through a circular tube filled with a UMF. The experimental data in- local thermal equilibrium (LTE) model. The temperature difference
dicated a significant improvement in the heat transfer rate at the cost of between the fluid and solid phases is taken into consideration in the
a pressure drop increase. Hamadouche et al. [18] found that inserting LTNE model, while it is neglected in the LTE model. In fact, the LTE
UMFs in a turbulent air flow improves the heat transfer by approxi- model is not suitable if there is obvious difference between the fluid and
mately 300% compared with an empty channel with adopting lower solid phase temperatures. The present author Xu [23] proposed a
velocity, which reduces the supplied power. Xia et al. [19] conducted mathematical model for fully developed forced convective heat transfer
the experimental measurement to investigate the heat transfer behavior in a tube partially filled with uniform metal foam by adopting the
of forced airflow in three kinds of Cu, Ni and SiC UMFs, and found that Brinkman-Darcy and the LTNE model. Lu et al. [24] analytically studied
the volumetric heat transfer coefficients decrease with an increase in the forced convection flow and heat transfer characteristics in parallel-
porosity and increase as pore density increases. Ghafarian et al. [20] plate channel partially filled with UMFs. More recently, the present
numerically conducted the fluid dynamics analysis of forced convective authors Xu et al. [25] studied forced convection in a mini/microchannel
heat transfer of oscillating flow through a channel filled with UMF. The filled with microfoam in the condition of uniform heat fluxes with LTNE
results showed significant heat transfer enhancements by inserting UMF effect. The results showed that, with the increased Knudsen number, the
in the channel and an increase in thermal conductivity of the metal friction factor gradually decreases and the decreasing amplitude in slip
foam and Reynolds number can significantly increase the heat transfer. flow region is very large. In the present study, to predict fluid and
Chen et al. [21] conducted a numerical investigation for enhanced heat thermal transport in the tubes sintered with partially-filled GMFs with
transfer from multiple discrete heated sources in a horizontal channel uniform heat flux boundary condition, the Brinkman-extended Darcy
by UMF porous layer. Both Darcy-Brinkman-Forchheimer flow model momentum model and LTNE model are adopted in GMF region and the
and two-equation energy model based on local thermal non-equilibrium momentum and energy conservation equations are used in GMF-free

102
Z.G. Xu et al. Applied Thermal Engineering 137 (2018) 101–111

region. Temperature profiles and velocity distributions can be predicted 1 ∂ ⎛ ∂〈Tf 〉 ⎞ ∂〈Tf 〉
k fe1 r −h sf α s1f (〈Ts 〉−〈Tf 〉) = ρf c f 〈u〉 .
and explicit expressions for the friction factor and Nusselt number can r ∂r ⎝ ∂r ⎠ ∂x (9)
be derived from the solution. The effects of gradients of GMF mor-
phology parameters, such as porosity, pore density and foam thickness
on flow resistance and heat transfer performance are investigated in the (3) GMF region II (rm < r < ro ):
present study. d〈p〉 μ 1 ∂ μ
− + f ⎛r ∂〈u〉 ⎞− f 〈u〉 = 0,
dx ε2 r ∂r ⎝ ∂r ⎠ K2 (10)
2. Problem description
1 ∂ ⎛ ∂〈Ts 〉 ⎞
k se2 r −hsf α s2f (〈Ts 〉−〈Tf 〉) = 0,
2.1. Physical problem r ∂r ⎝ ∂r ⎠ (11)

The schematic diagram of the tube is shown in Fig. 1. As an ideal 1 ∂ ⎛ ∂〈Tf 〉 ⎞ ∂〈Tf 〉
k fe2 r −hsf α s2f (〈Ts 〉−〈Tf 〉) = ρf c f 〈u〉 .
heat transfer enhancement surface, annular GMFs with dense ligaments r ∂r ⎝ ∂r ⎠ ∂x (12)
are sintered into the inner wall of the tubes. The pore densities of
gradient metal foams are less than 40 PPI, but the porosities outstrip The energy balance equation can be obtained as follows by con-
0.8. The interface between the two foam layers is concentric. sidering uniform heat flux boundary condition.

ρf c f um ·πro2·d〈Tfb 〉 = qw ·2πro·dx (13)


2.2. Mathematical description for forced convection heat transfer in a single
tube sintered with partially-filled gradient metal foams Then Eqs. (6), (9)and (12) can be rewritten as:
1 ∂ ∂〈Tf 〉 2q 〈u〉
To further describe the physical problem, the fluid flow in the tubes kf (r )= w ,
r ∂r ∂r ro um (14)
is assumed to be incompressible and laminar, and thermo-physical
properties in the present study is temperature independent. Wall heat 1 ∂ ⎛ ∂〈Tf 〉 ⎞ 2q 〈u〉
flux distribution imposed on the outer tube wall is assumed to be uni- k fe1 r −hsf α s1f (〈Ts 〉−〈Tf 〉) = w ,
r ∂r ⎝ ∂r ⎠ ro um (15)
form, and metal foam layers are homogeneous and isotropic porous
media. Because GMFs are sintered on the inner tube surface, contact 1 ∂ ⎛ ∂〈Tf 〉 ⎞ 2q 〈u〉
k fe2 r −hsf α s2f (〈Ts 〉−〈Tf 〉) = w .
thermal resistance between solid wall and foam layer can be neglected. r ∂r ⎝ ∂r ⎠ ro um (16)
The entrance effect for GMFs, natural convection thermal radiation,
gravity effects, tube wall thickness, and heat conduction in flow di- As indicated in Appendix A, the values of the parameters K1, K2 (the
rection are assumed to be neglected. With these assumptions, the permeabilities of metal foam layer I and metal foam layer II (as shown
equations in the GMF-free region and GMF region can be obtained: in Fig. 1), k e1,k e2 (the effective thermal conductivities of metal foam
layer I and metal foam layer II), hsf (local heat transfer coefficient), and
∂〈u〉 α s1f , α s2f (the specific surface areas of metal foam layer I and metal foam
= 0, 〈v〉 = 0, 〈p〉 = constant,
∂x (1) layer II) can be obtained from the literatures [26–28]. The solid effec-
tive thermal conductivities of metal foam layer I and metal foam layer
∂ ⎛ 〈Tf 〉−〈Tw 〉 ⎞ ∂ ⎛ 〈Ts1 〉−〈Tw 〉 ⎞ ∂ ⎛ 〈Ts2 〉−〈Tw 〉 ⎞
⎜ ⎟ = 0, ⎜ ⎟ = 0, ⎜ ⎟ = 0,
∂x ⎝ 〈Tfb 〉−〈Tw 〉 ⎠ ∂x ⎝ 〈Tfb 〉−〈Tw 〉 ⎠ ∂x ⎝ 〈Tfb 〉−〈Tw 〉 ⎠
(2)
d〈Tfb 〉 d〈Tw 〉
= = constant, h x = constant.
dx dx (3)
Then the following equations are obtained.
d〈Tfb 〉 d〈Tw 〉 ∂〈Tf 〉 ∂〈Ts1 〉 ∂〈Ts2 〉
= = = =
dx dx ∂x ∂x ∂x (4)
In the GMF region, the Brinkman-extended Darcy model is used for
momentum equations, the two-equation model considering local
thermal non-equilibrium (LTNE) effect is used for the energy equations.
The equations governing fluid flow and heat transfer in a tube sintered
with partially-filled GMFs can be can be greatly simplified as follows: (a) Cross section
(1) GMF-free region (0 < r < ri ):
d〈p〉 1 ∂ ⎛ ∂〈u〉 ⎞
− + μf r = 0,
dx r ∂r ⎝ ∂r ⎠ (5)

1 ∂ ∂〈Tf 〉 ∂〈Tf 〉
kf (r ) = ρf c f 〈u〉 .
r ∂r ∂r ∂x (6)

(2) GMF region I (ri < r < rm ):


d〈p〉 μ 1 ∂ μ
− + f ⎛r ∂〈u〉 ⎞− f 〈u〉 = 0,
dx ε1 r ∂r ⎝ ∂r ⎠ K1 (7)

1 ∂ ⎛ ∂〈Ts 〉 ⎞
k se1 r −hsf α s1f (〈Ts 〉−〈Tf 〉) = 0,
r ∂r ⎝ ∂r ⎠ (8) (b) Axial section
Fig. 1. The tube partially sintered with gradient foams.

103
Z.G. Xu et al. Applied Thermal Engineering 137 (2018) 101–111

II, k se1, k se2 can be obtained from the equations k se1 = k e1 |kf = 0 , ∂ 2U ∂U 2 2
R2 +R −s2 R (U + P2) = 0,
k se2 = k e2 |kf = 0 , respectively, while the liquid effective thermal con- ∂R2 ∂R (27)
ductivities, k fe1, k fe2 can be obtained from the expressions k fe1 = k e1 |ks1= 0 ,
∂2θf 1 ∂θf ⎞
k fe2 = k e2 |ks2= 0 , respectively. C2 ⎛⎜ 2 + ⎟ + D 2 (θs + θf ) = 2U ,
The dimensional boundary conditions are as follows: ⎝ ∂R R ∂R ⎠ (28)
∂〈u〉 ∂〈Tf 〉 ∂2θs 1 ∂θs
r = 0: = 0, = 0, + −D2 (θs−θf ) = 0.
∂r ∂r (17) ∂R2 R ∂R (29)
∂〈u〉 ⎞ μ f ∂〈u〉 The dimensionless boundary conditions are as follows:
r = ri: 〈u〉|ri− = 〈u〉|ri+ , μ f ⎛ = ⎛ ⎞ , 〈Tf 〉|ri− = 〈Tf 〉|ri+ ,
⎝ ∂r ⎠ ri
− ε1 ⎝ ∂r ⎠ +
ri ∂U ∂θf
R = 0: = 0, = 0,
∂r ∂r (30)
∂〈Ts1 〉 ⎞ ∂〈Tf 〉 ⎞
k se1 ⎛
⎜ = kf ⎛

⎝ ∂r ⎠ ⎝ ∂r ⎠ − ∂U ∂U
ri R = Ri : U |Ri− = U |Ri+ , ε1 = , θf |Ri− = θf |Ri+ ,
∂R −
Ri ∂R Ri
+
∂〈Ts1 〉 ⎞ ∂〈Tf 〉 ⎞ ⎞
= h s1f α s1f (〈Ts1 〉−〈Tf 〉) ⎜⎛k se1 ⎛ + k fe1 ⎛
⎜ ⎟ ⎟

⎝ ⎝ ∂r ⎠ ⎝ ∂r ⎠ ⎠ + ∂θs ∂θ ∂θf ⎞ ⎞ ∂θf ⎞


ri
= A1 (θs−θf ) ⎜⎛ ⎛ s ⎞ + C1 ⎛ ⎟ ⎜ ⎟ = B1 ⎛ ⎜ ⎟

(18) ∂R ⎝⎝ ∂R ⎠ ⎝ ∂R ⎠ ⎠ + ⎝ ∂R ⎠ −
Ri Ri (31)
μ f ∂〈u〉 − μ f ∂〈u〉
r = rm: 〈u〉|rm− = 〈u〉|rm+ , ⎛ ⎞ rm = ⎛ ⎞ , 〈Tf 〉|rm− ∂U ⎞ ∂U ⎞
ε1 ⎝ ∂r ⎠ ε2 ⎝ ∂r ⎠ + R = Rm : U |Rm− = U |Rm+ , ε2 ⎛ = ε1 ⎛ , θf |Rm−
rm ⎝ ∂R ⎠ −
Rm ⎝ ∂R ⎠ Rm
+

∂〈Ts 〉 ⎞ ∂〈Ts 〉 ⎞ ∂θ ∂θ
= 〈Tf 〉|rm+ , k se1 ⎛ = k se2 ⎛ = θf |Rm+ , C2 ⎛ s ⎞ = C1 ⎛ s ⎞
⎝ ∂r ⎠ −
rm ⎝ ∂r ⎠ +
rm (19) ⎝ ∂R ⎠ − ⎝ ∂R ⎠ +
(32)
Rm Rm

r = ro: 〈u〉 = 0, 〈Tf 〉 = 〈Tw 〉, 〈Ts 〉 = 〈Tw 〉. (20) R = Ro : U = 0, θf = θw, θs = θw. (33)
The dimensionless parameters are as follows:
〈T 〉−〈Tw 〉 〈T 〉−〈Tw 〉 〈T 〉−〈Tw 〉 r 〈u〉 2.3. Analytical solutions for velocity and temperature
θ= , θs = s , θf = f ,R= ,U= ,P
qw ro/ k se qw ro/ k se qw ro/ k se ro um
K1 + K2 d〈p〉 The velocity distribution and temperature profile functions (34) and
= , (35) can be derived from Eqs. (22) and (23) by considering boundary
2μ f um dx
conditions (30) and (31).
K1 d〈p〉 K2 d〈p〉 K1 + K2 K1 K2
P1 = , P2 = , Da = , Da1 = , Da2 = , A1
μ f um dx μ f um dx 2ro2 ro2 ro2 (1) GMF-free region (0 ⩽ R ⩽ Ri ):
hsf ro hsf ro
= , A2 = ,
k se1 k se2 R2
k fe1 k fe2
U = P⎛ ⎜ + E0 ⎞ ⎟

B=
kf
, B1 =
kf
, B2 =
kf
, C1 = , C2 = , D1 4
⎝ Da ⎠ (34)
(k se1 + k se2) / 2 k se1 k se2 k se1 k se2

hsf αs1f ro2 hsf αs2 f ro2 P R4 E


= , D2 = , θf = [ + 0 R2 + H0]
k se1 k se2 B 32Da 2 (35)

ε1 ε2 D1 (C1 + 1) D2 (C2 + 1) The velocity distribution and temperature profile functions (36)-
s1 = , s2 = , t1 = , t2 = . (41) can be derived from Eqs. (24) and (29) by considering boundary
Da1 Da2 C1 C2 (21)
conditions (31)–(33).
Therefore, the dimensionless governing equations can be derived as
follows: (2) GMF region I (Ri ⩽ R ⩽ Rm ):
U = P1 [E1 I0 (sR) + E2 K 0 (sR)−1] (36)
(1) GMF-free region (0 < R < Ri ):
D
1 ∂ ⎛ ∂U ⎞ P (C1 − 1) ⎛1 − 21 ⎞
R = , 2P1 ⎧ ⎝ s ⎠
R ∂R ⎝ ∂R ⎠ Da (22) θf = −
C1 + 1 ⎨ C1 s 2 − D1 (C1 + 1)
[E1 I0 (sR) + E2 I0 (sR)] + (H1 I0 (sR) + H2 I0 (sR))

∂2θf 1 ∂θf ⎞
B ⎛⎜ 2 + ⎟ = 2U . R2 R4 E0 Ri2 ⎫
∂R R ∂R ⎠ (23) − i
+ ⎛ 16Da + ⎞ lnR + 1 − 12 + 1
⎝ 4
⎝ 1 2
⎠ 4 s D1 (C1 + 1) ⎬

(2) GMF region I (Ri < R < Rm ): (37)

∂ 2U ∂U 2 2 2P1 ⎧ D1 (C1 + 1)
R2 +R −s1 R (U + P1) = 0, θs = − 2
[E1 I0 (sR) + E2 I0 (sR)]
∂R2 ∂R (24) C1 + 1 ⎨
⎩ C1 s −D1 (C1 + 1)
∂2θf 1 ∂θf ⎞ R2 R4 E0 Ri2 ⎞ 1 1
C1 ⎛⎜ 2 + ⎟ + D1 (θs + θf ) = 2U ,
−(H1 I0 (sR) + H2 I0 (sR))− + ⎛⎜ i + ⎟ lnR + −
∂R R ∂R ⎠ 4 ⎝ 16Da1 2 ⎠ 4 s2
⎝ (25)
1 ⎫
∂2θs 1 ∂θs −
+ −D1 (θs−θf ) = 0. D1 (C1 + 1) ⎬
⎭ (38)
∂R2 R ∂R (26)

(3) GMF region II (Rm < R < 1): (3) GMF region II (Rm ⩽ R ⩽ 1):
U = P2 [E1 I0 (sR) + E2 G0 (sR)−1] (39)

104
Z.G. Xu et al. Applied Thermal Engineering 137 (2018) 101–111

D than that of 0.98 & 0.9. The metal skeleton attached to the inner tube
2P2 ⎧ (C2−1)(1− 22 )
s
θf = − [E1 I0 (sR) + E2 I0 (sR)] wall of the GMF 0.98 & 0.8 is thicker than GMF 0.98 & 0.9, then the
C2 + 1 ⎨ C2 s 2−D2 (C2 + 1)
⎩ former’s flow resistance is higher and its liquid volume ratio is lower.
R2 R4 E ′R 2 1 1 Fig. 4(a) shows that, the maximum velocity the GMF with porosity
+ (H1 I0 (sR) + H2 I0 (sR))− + ⎛⎜ i + 0 i ⎞⎟ lnR + − 2 gradient 0.98 & 0.8 is higher than that of 0.8 & 0.98. In the foam region,
4 ⎝ 16Da 2 2 ⎠ 4 s
the liquid volume ratios of the two GMFs are the same. However, the
1 ⎫ flow resistance of the inner foam layer of the GMF 0.98 & 0.8 is lower
+
D2 (C2 + 1) ⎬ than the 0.98 & 0.8, the central liquid flow state in the GMF-free region
⎭ (40)
is less affected by the solid-liquid interface, which leads to higher liquid
2P2 ⎧ D2 (C2 + 1) velocity in the tube central region.
θs = − 2−D (C + 1)
[E1 I0 (sR) + E2 I0 (sR)] Fig. 4(b) presents pore density gradient effect on velocity distribu-
C2 + 1 ⎨
⎩ C2 s 2 2
tions at the fixed porosity of 0.9. The results show that, in the GMF
R2 R4 E ′R 2 1 1 region, velocity in the tubes partially sintered with the gradient copper
−(H1 I0 (sR) + H2 I0 (sR))− + ⎛⎜ i + 0 i ⎞⎟ lnR + − 2
4 ⎝ 16Da2 2 ⎠ 4 s foams with the pore density gradients 5 PPI & 20 PPI and 5 PPI & 40 PPI
increases from the inner wall to the tube center because the flow re-
1 ⎫
− sistance decreases with decreasing of pore density. The results also
D2 (C2 + 1) ⎬
⎭ (41) show that, in the GMF region, velocity in the tubes partially sintered
The friction factor (f) and overall Nusselt number (Nu) are defined with the gradient copper foams with the pore density gradient 20 PPI &
as: 5 PPI firstly decreases then increases from the inner wall to the tube
center. The mean velocity in the GMF-free region is much higher than
−dp /dx·2r0 −8P
f= = , that in the GMF region, which implies that the velocity in the foam
1/2ρf u m2 Da·Re (42) region is very small for a big frontal velocity [23]. Fig. 4(a) shows that,
h·2ro 2 in the GMF-free region, the maximum velocity of the gradient copper
Nu = =− . foam with the pore density gradient 5 PPI & 40 PPI is higher than that
kf B·θfb (43)
of 5 PPI & 20 PPI. The metal skeleton attached to the inner tube wall of
where θfb can be obtained from Appendix A. the GMF 5 PPI & 40 PPI is denser than 5 PPI & 20 PPI, thus the flow
j
In the present study, the parameter 1/3 , involving heat transfer and resistance of the former is higher. However, the liquid volume ratios of
f
flow resistance is used to evaluate the comprehensive performance of the two GMFs are the same. The extrusion effect on the central liquid by
the tubes sintered with partially-filled GMFs: the GMF 5 PPI & 40 PPI is more obvious. Fig. 4(a) shows that, the
maximum velocity the gradient copper foam with porosity gradient 5
j Nu
= PPI & 20 PPI is higher than that of 20 PPI & 5 PPI. In the foam region,
f 1/3 Re ·(Pr·f )1/3 (44) the liquid volume ratios of the two GMFs are the same. However, the
where j is the dimensionless Colburn factor for heat transfer. flow resistance of the inner foam layer of the GMF 5 PPI & 20 PPI is
lower than the 20 PPI & 5 PPI, the central liquid flow state in the GMF-
Nu
j= free region is less affected by the solid-liquid interface, which leads to
Re ·Pr 1/3 (45)
higher liquid velocity in the central region of the tube.
Fig. 4(c) shows velocity distributions of the GMF with the pore
3. Results and discussion density gradient 20 PPI & 5 PPI for different dimensionless interfacial
radii. The results indicated that, in the GMF region, the velocity de-
3.1. Validation of solution creases with increasing interfacial radius or decreasing GMF volume
ratio. The peak value exists for that of the tube partially sintered with
The grid independence are checked, as shown in Fig. 2. In the GMF because the GMF-free region is much more permeable than the
present study, the final adopted grid number is 192 × 82. The numer- GMF region. The peak velocity firstly increases and then decreases with
ical iteration can be treated as convergent when relative variations in increasing radius. When the value of interfacial radius is high, the tube
pressure drop and Nusselt number between successive iterations are partially filled with less GMF shows a relatively uniform profile. This is
less than 10−7. The results of the present solution are compared with because the drag effect by the GFM skeletons and the tube inner wall
the known results of the tube partially filled with UFMs. By setting the
second foam layer pore density as 10 PPI, which is the same with the
650
first foam layer, the present solution coincides with the solution in
UMF-filled tube [23], as shown in Fig. 3, and its feasibility can be va- 192x82 202x82
lidated 182x82
162x82
3.2. Velocity distributions and temperature profiles 172x72
600
The effects of key parameters on the velocity distributions are pre-
Nu

sented in Fig. 4(a)–(c). Fig. 4(a) shows the porosity gradient effect on
velocity profile at the fixed pore density of 20 PPI. The results show
that, in the GMF region, velocity in the tubes partially sintered with the 152x62
gradient copper foams with the porosity gradients 0.98 & 0.8 and 0.98 550
& 0.9 increases from the inner wall to the tube center because the flow
resistance decreases with increasing of porosity. The results also show 132x52
that, in the GMF region, velocity in the tubes sintered with the partially-
filled gradient copper foams with the porosity gradient 0.8 & 0.98 1 2 3 4 5 6 7 8
firstly decreases then increases from the inner wall to the tube center. Grid No.
Fig. 4(a) shows that, in the GMF-free region, the maximum velocity of
the gradient copper foam with the porosity gradient 0.98 & 0.8 is higher Fig. 2. Nu number variation with grid number for Re = 1600.

105
Z.G. Xu et al. Applied Thermal Engineering 137 (2018) 101–111

200

Dimensionless velocity U
ω =20PPI εg=0.8&0.98
150 Ri=0.1
εg=0.98&0.8
r0=0.01m
Re=1600
εg=0.98&0.9
100
water/copper

50

0
-1.0 -0.5 0.0 0.5 1.0
Dimensionless radius R
Fig. 3. Validation of the present solution.
(a) Porosity gradient effects on velocity profile

decreases.
Morphology parameter gradient effects on temperature profiles are
200 ωg=5PPI&20PPI
180
presented in Fig. 5. Fig. 5(a) shows pore density gradient effect on the ωg=20PPI&5PPI

Dimensionless velocity U
temperature profile. The excess solid temperature of the GMF 5 PPI & 160 Ri=0.1
20 PPI is higher than 5 PPI & 40 PPI because thicker fibers results in less 140 r0=0.01m ωg=5PPI&40PPI
thermal resistance in the foam matrix when the porosity is fixed. The Re=1600
120
fluid excess temperature of the GMF 5 PPI & 20 PPI is higher than 5 PPI water/copper
& 40 PPI in the whole region. The heat transfer coefficient of the tube 100
with the GMF 5 PPI & 40PPI is higher due to the higher surface area in 80
the foam region. Fig. 5(a) shows that, for the tube partially with GMF 5 60
PPI & 20 PPI, the solid temperature exceeds the liquid temperature in
40
the whole foam region. The liquid velocity gradually increases from the
inner wall to the tube central region (as shown in Fig. 4(b)), and the 20
skeleton heat can be transfer to liquid without delay. Fig. 5(a) shows 0
that, for the tube partially with GMF 5 PPI & 40 PPI, the solid tem- -1.0 -0.5 0.0 0.5 1.0
perature exceeds the liquid temperature in the 5 PPI foam region while Dimensionless radius R
the solid temperature is approximately equal to the liquid temperature (b) Pore density gradient effects on velocity profile
in the 40 PPI foam region. The liquid velocity increases sharply from
the 40 PPI foam region to the 5 PPI foam region (as shown in Fig. 4(b)),
which means the skeleton heat can be transfer to liquid without delay 100
Ri=0.1 ωg=20PPI&5PPI
inside 5 PPI while heat transfer between the skeletons and liquid is
Dimensionless velocity U

dominated by thermal conduction inside 40 PPI foam. Fig. 5(a) shows Ri=0.3 ε=0.9
that, heat transfer coefficient of the GMF 5 PPI & 40 PPI is higher than Ri=0.5 r0=0.01m
GMF 5 PPI & 20 PPI because of more solid-liquid surface area.The re- Ri=0.7 Re=1600
sults of Fig. 5(a) also show that, the excess solid temperature of the water/copper
GMF 20 PPI & 5 PPI is higher than 5 PPI & 20 PPI due to the higher heat 50
conduction capacity of the metal skeleton close to the inner tube wall.
The excess fluid temperature of the GMF 5 PPI & 20 PPI is higher than
20 PPI & 5 PPI in the foam region0.88 < |R| < 1,while the trend con-
verses in the foam region |R| < 0.88. Fig. 5(a) shows that, even with the
same solid-liquid surface area, heat transfer coefficient of GMF 5 PPI &
20 PPI is higher than GMF 20 PPI & 5 PPI because of higher mass flow 0
-1.0 -0.5 0.0 0.5 1.0
rate. The results of Fig. 5(a) also show that, for the tube partially with
GMF 20 PPI & 5 PPI, the solid temperature approximately equal to the Dimensionless radius R
liquid temperature in the foam region0.25 < |R| < 0.55,while the solid (c) Interfacial radius effects on velocity profile
temperature exceeds the liquid temperature on the other foam region.
For the tube partially with GMF 5 PPI & 20 PPI, the solid temperature Fig. 4. The effects of key parameters on velocity distribution.
approximately equal to the liquid temperature in the foam region
0.7 < |R| < 0.85, while the solid temperature exceeds the liquid tem- the two foam layer interface may swirl because of the drag effect of the
perature in the other foam region.Fig. 5(b) shows porosity gradient metal skeletons on the interface. The liquid temperature is approxi-
effect on the temperature profile. The excess solid and fluid tempera- mately equal to the solid temperature in the foam re-
ture of the GMF 0.98 & 0.8 is higher than 0.8 & 0.98 due to the higher gion0.62 < |R| < 0.9,while the solid temperature exceeds the liquid
heat conduction capacity of the metal skeleton close to the inner tube temperature in the other foam region. For the tube partially with GMF
wall. The heat transfer coefficient of the tube with the GMF 0.8 & 0.98 0.98 & 0.8, the liquid temperature is approximately equal to the solid
is higher than GMF 0.98 & 0.8. For the tube partially with GMF 0.8 & temperature in the foam region 0.22 < |R| < 0.46 while the solid tem-
0.98, the abnormal phenomenon exists in the foam region perature exceeds the liquid temperature in the other foam region.
0.48 < |R| < 0.62 because the liquid temperature exceeds the solid
temperature. When the porosity decreases sharply, the liquid close to

106
Z.G. Xu et al. Applied Thermal Engineering 137 (2018) 101–111

inner tube wall. The friction factor of the GMF 0.98 & 0.8 is higher than
0.0
0.98 & 0.9 because of the higher solid volume ratio.
ωg=5PPI&20PPI The dependence of the friction factor on the pore density gradient
-1.0x10
-4 for different pore densities is shown in Fig. 6(b). With the fixed pore
density gradients of 5 PPI & 10 PPI and 20 PPI & 5 PPI, the friction
factor firstly increases slightly and then considerably decreases with
(T-Tw)/qw

-4
-2.0x10
ωg=5PPI&40PPI increasing porosity. However, when the pore density gradients are 5
ε=0.9 PPI & 20 PPI and 5 PPI & 40 PPI, increasing porosity decreases the
-3.0x10
-4 Re=1600 liquid friction factor in the whole porosity region.
Ri=0.1 solid The pore density gradient orders enhancing friction factor are 20
water vs.Cu liquid
PPI & 5 PPI, 5 PPI & 10 PPI, 5 PPI & 20 PPI and 5 PPI & 40 PPI. The
-4.0x10
-4 solid
friction factor of the gradient 5 PPI & 40 PPI is higher than 5 PPI & 20
PPI and 5 PPI & 10 PPI because of the denser skeleton. Particularly,
-5.0x10
-4 even with the same combination of 5 PPI and 20 PPI, the friction factor
-1.0 -0.5 0.0 0.5 1.0 of the gradient 5 PPI & 20 PPI is higher than 20 PPI & 5 PPI due to
Dimensionless radius R serious drag effect by the denser metal skeletons close to the inner tube
0.0 wall.
ωg=20PPI&5PPI
3.4. Heat transfer performance
-4
-1.0x10
ωg=5PPI&20PPI
(T-Tw)/qw

In the GMF region, the total thermal resistance includes the con-
ductive thermal resistance of the metal skeletons, the conductive
ε=0.9 liquid
-4 thermal resistance of the liquid and the local convective thermal re-
-2.0x10 Re=1600 solid
Ri=0.1
sistance on the solid-liquid contacting surface.
liquid
Morphology parameter effects on Nusselt number are shown in
water/Cu solid
Fig. 7. Pore density gradient effects on Nusselt number for different
-4
-3.0x10 porosities are shown in Fig. 7(a). For a fixed pore density gradient,
Nusselt number decreases considerably with increasing porosity. Metal
-1.0 -0.5 0.0 0.5 1.0
Dimensionless radius R
(a) Pore density gradient effects on temperature profile
0.0

εg=0.98&0.8
-4
-1.0x10
εg=0.8&0.98
(T-Tw)/qw

-4
ω=20 PPI
-2.0x10 Re=1600 liquid
Ri=0.1 solid
liquid
water/Cu solid
-4
-3.0x10

-1.0 -0.5 0.0 0.5 1.0


Dimensionless radius R
(b) Porosity gradient effects on temperature profile
Fig. 5. Morphology parameter gradient effects on temperature profile.

3.3. Flow characteristics

Fig. 6 shows morphology parameter effects on the friction factor


with the tube diameter of 0.01 m and Reynolds number of 1600. The
dependence of the friction factor on the porosity gradient for different
pore densities is shown in Fig. 6(a). With the fixed porosity gradient,
the friction factor increases considerably with increasing pore density
because the stirring extent of the obstructing solid is increased by the
denser and denser skeletons. At the low pore density, the porosity
gradient has slight effect on the friction factor while the effect increases
with increasing pore density. Fig. 6 shows that, even with same surface
area, the friction factor of the GMF 0.98 & 0.8 is higher than 0.8 & 0.98
due to serious drag effect by the thicker metal skeletons close to the Fig. 6. Morphology parameter gradient effects on flow resistance.

107
Z.G. Xu et al. Applied Thermal Engineering 137 (2018) 101–111

lineally increases with increasing Reynolds number.


Fig. 9 presents thermal conductivity ratio effect on Nusselt number
for the two reversal GMFs, 20 PPI & 5 PPI and 5 PPI & 20 PPI, with the
fixed fluid of water. It is can be found that, for the GMF, thermal
conductivity ratio has a significant influence on Nusselt number. When
k f / k s > 0.01, Nusselt number decreases mildly with increasing thermal
conductivity ratio and approaches a constant value which is much more
than 4.36, which is the value of classical solution of laminar and fully
developed flow in a hollow tube with uniform surface-heat flux con-
dition. This implies that thermal resistance of heat conduction accounts
for the majority of the total thermal resistance and that heat transfer is
enhanced. When k f / k s < 0.01, Nusselt number increases sharply with
decreasing thermal conductivity ratio and approaches a high constant
value and the heat transfer is enhanced further correspondingly. This is
because the thermal resistance of heat conduction is negligible and
local convective thermal resistance dominates the entire heat transfer
process [23].Interfacial radius effect on Nusselt number for different
radii of the tube partially sintered with GMF with the pore gradient 20
PPI & 5 PPI at the porosity 0.9 is shown in Fig. 10. Nusselt number does
not monotonically vary with increasing interfacial radius. Due to the
sharp variation of flow-permeability from the foam region across the
porous-fluid interface to the fluid region, an increase in interfacial ra-
dius leads to the liquid velocity in the GMF region being reduced and
local convective thermal resistance increased. On the other hand, an
increase in interfacial radius leads to solid-liquid surface area being
decreased. Thus, total convective thermal resistance increased. Hence,
the heat conduction becomes the dominated factor of heat transfer in
the high value of interfacial radius. Fig. 10 shows that, Nusselt number
increases sharply with decreasing interfacial radius in the range
0.1 < Ri < 0.7. When interfacial radius increases, the liquid velocity in
the GMF region increases considerably (as shown in Fig. 4) and solid-
liquid surface area increases, which leads to total convective thermal
resistance being decreased. Moreover, the increasing volume of metal
skeleton reduces solid thermal resistance. Therefore, the total thermal
resistance decreases sharply with decreasing interfacial radius.Fig. 10
also shows that, Nusselt number decreases with increasing tube dia-
meter in the range 0.35 < Ri < 0.5. Nusselt number increases with in-
creasing tube diameter in the range Ri < 0.35, but the effect is not
significant. Nusselt number increases significantly with increasing tube
Fig. 7. Morphology parameter gradient effects on Nu number. diameter when Ri > 0.7 . Fig. 11 shows the interfacial radius effect on
the comprehensive performance parameter j/ f 1/3 with the two Rey-
fiber diameter decreases with increasing porosity, which leads to the nolds numbers of 800 and 1600. It is found that j/ f 1/3 decreases sharply
increasing of the conductive thermal resistance of the metal skeletons. with increasing interfacial radius in the rangeRi < 0.3. However, j/ f 1/3
Fig. 7 (a) also shows that, Nusselt number decreases with the order of 5 firstly decreases and then increases mildly with increasing interfacial
PPI & 10 PPI, 5 PPI & 20 PPI and 5 PPI & 40 PPI. Although the solid- radius in the range Ri > 0.3.4. Conclusions
liquid surface area increases with the order of 5 PPI & 10 PPI, 5 PPI & Forced convective heat transfer in tubes sintered with partially-
20 PPI and 5 PPI & 40 PPI, the flow resistance increases considerably.
The two opposite factors leads to the results. Fig. 7 (a) also shows that,
Nusselt number of the GMF 20 PPI & 5 PPI is higher than 5 PPI & 20 PPI
due to the lower friction factor (as shown in Fig. 6 (b)).
Porosity gradient effects on Nusselt number for different pore den-
sities are shown in Fig. 7(b). For a fixed porosity gradient, as pore
density increases, Nusselt number decreases sharply. Nusselt number
decreases with the order of 0.9 & 0.8, 0.98 & 0.8 and 0.98 & 0.9 because
solid-liquid surface area decreases. Fig. 7 (b) also shows that, Nusselt
number of the GMF 0.98 & 0.8 is higher than GMF 0.8 & 0.98.
Reynolds number effect on Nusselt number for different pore den-
sity gradients is shown in Fig. 8. Nusselt number increases with Rey-
nolds number increasing, which is ascribed to the fact that the increase
in Reynolds number can lead to the decreased local convective thermal
resistance in the foam region. Undoubtedly, increasing Re can improve
the heat transfer performance [23]. For the GMF 20 PPI & 5 PPI, Nusselt
number increases firstly sharply and then mildly with increasing Re.
The maximum Nusselt number approaches the constant value of 600.
For the GMF 5 PPI & 20 PPI and 5 PPI & 40 PPI, Nusselt number almost
Fig. 8. Re number effect on Nu number.

108
Z.G. Xu et al. Applied Thermal Engineering 137 (2018) 101–111

the LTNE effect with the uniform heat flux boundary condition. The
main conclusions are as follows:

(1) In the GMF region, velocity in the tubes partially sintered with the
gradient copper foams with the porosity gradients 0.98 & 0.8 and
0.98 & 0.9 or the pore density gradients 5 PPI & 20 PPI and 5 PPI &
40 PPI increases from the inner wall to the tube center because the
flow resistance decreases with increasing porosity or decreasing
pore density. The velocity decreases with increasing interfacial ra-
dius.
(2) The excess solid temperature of the GMF 5 PPI & 20 PPI is higher
than 5 PPI & 40 PPI because thicker fibers results in less thermal
resistance in the foam matrix when the porosity is fixed. The fluid
excess temperature of the GMF 5 PPI & 20 PPI is higher than 5 PPI &
40 PPI in the whole region. The excess solid temperature of the
GMF 20 PPI & 5 PPI is higher than 5 PPI & 20 PPI. The excess fluid
temperature of the GMF 5 PPI & 20 PPI is higher than 20 PPI & 5
Fig. 9. Thermal conductivity ratio on Nu number. PPI in the foam region 0.88 < |R| < 1, while the trend converses
|R| < 0.88. The excess solid and fluid temperature of the GMF 0.98 &
0.8 is higher than 0.8 & 0.98.
(3) Porosity gradient has slight effect on the friction factor at a low
pore density, while the effect increases with increasing pore den-
sity. The friction factor of the GMF 0.98 & 0.8 is higher than GMF
0.8 & 0.98. With the fixed pore density gradients of 5 PPI & 10 PPI
and 20 PPI & 5 PPI, the friction factor firstly increase slightly and
then decreases considerably with increasing porosity. However,
when pore density gradients are 5 PPI & 20 PPI and 5 PPI & 40 PPI,
the friction factor decreases with increasing porosity in the whole
porosity region.
(4) Nusselt number decreases with the pore density gradient order of 5
PPI & 10 PPI, 5 PPI & 20 PPI and 5 PPI & 40 PPI. Nusselt number of
the GMF 20 PPI &5 PPI is higher than 5 PPI & 20 PPI. Nusselt
number decreases with the order of 0.9 & 0.8, 0.98 & 0.8 and 0.98 &
0.9. Nusselt number of the GMF 0.98 & 0.8 is higher than 0.8 &
0.98. For a fixed pore density gradient, Nusselt number increases
with increasing Reynolds number. For the tube partially sintered
with GMF with the pore gradient 20 PPI & 5 PPI at the porosity 0.9,
Fig. 10. Interfacial radius on Nu number.
Nusselt number increases sharply with decreasing interfacial radius
in the range 0.1 < Ri < 0.7 .

Acknowledgements

This work is supported by the National Natural Science Foundation


of China (Grant No. 51576126).

Fig. 11. Interfacial radius effect on comprehensive performance.

filled gradient metal foams is numerically investigated by considering

Appendix A

The parameters used in the present study are as followings:

1. Pore diameter (dp ) and fiber diameter (df ) [26]

109
Z.G. Xu et al. Applied Thermal Engineering 137 (2018) 101–111

dp = 0.0254/ ω,
df
dp
= 1.18 1−ε
3π ( 1
1 − e−(1 − ε )/0.04 )
2. Permeability of metal foams (K ) [26]
K = 0.00073(1−ε )−0.224 (df / dp)−1.11dp2

3. Surface area density of metal foams(K ) [27]


3πdf [1−e−(1 − ε )/0.04]
αsf =
(0.59dp)2

4. Local convective heat transfer coefficients in tubes [8]


0.4 0.37
⎧ 0.76Red Pr kf / df ,(1 ⩽ Red⩽40)
⎪ 0.5 0.37 3
hsf = 0.52Red Pr kf /df ,(40 ⩽ Red⩽10 )
⎨ 0.6 0.37 3 5
⎪ 0.26Red Pr kf /df ,(10 ⩽ Red⩽10 )

5. Effective thermal conductivity [28]


1
ke = 2 (RA + RB + RC + RD )
,

RA =
(2e 2 + πλ (1 − e )) ks + (4 − 2e 2 − πλ (1 − e )) kf
(e − 2λ )2
RB =
(e − 2λ ) e 2ks + (2e − 4λ − (e − 2λ ) e 2) kf
( 2 − 2e )2
RC =
2πλ2 (1 − 2e 2 ) ks + 2( 2 − 2e − πλ2 (1 − 2e 2 )) kf
2e
RD =
e 2ks + (4 − e 2) kf

where,
2 (2 − (5 / 8) e3 2 − 2ε )
λ= π (3 − 4e 2 − e )
,e = 0.339

6. The constant used for formula (34), (35)

i R2
E0 = E1 I0 (sRi ) + E2 Go (sRi )− 4Da −1
where,
R
G1 (sRi) + s· i ·Go (s )
2
E1 = I0 (s ) G1 (sRi) + I1 (sRi) G0 (s )
,
R
I1 (sRi) − s· i ·Io (s )
2
E2 = I0 (s ) G1 (sRi) + I1 (sRi) G0 (s )

1
P=
2 Ri4
[E1 (I1 (s )−I1 (sRi ) Ri ) + E1 (G1 (s )−G1 (sRi ) Ri )] + + (E0 + 1) Ri2−1
s2 8Da

7. The constant used for formula (36)–(41)

1 ⎡ Ri2 R4 E0 Ri2
H0 = (E1 I0 (tRi ) + E2 G0 (tRi ))− i
+ ⎛ 16Da + ⎞ ln(Ri ) + 1 − 12 + 1 ⎤
C+1 4 2 4 s D (C + 1)
⎣ ⎝ ⎠ ⎦
G0 (t ) N3 − N2 N4
H1 = G0 (t ) N1 − I0 (t ) N2
N1 N4 − I0 (t ) N3
H2 = G0 (t ) N1 − I0 (t ) N2

where
DC
N1 = AI0 (tRi )− I1 (tRi )
C+1
DC
N2 = AG0 (tRi )− G1 (tRi )
C+1
Ri3 Ri Cs 2 A A R2
N3 = − 16(C + 1) Da + ⎡ −A0 ⎤− D (C + 1) + ⎛ i + A0 + 1⎞
2(C + 1) ⎣ Cs 2 − D (C + 1) ⎦ Cs 2 − D (C + 1) ⎝ 4Da ⎠
Cs 2
N4 =
Cs 2 − D (C + 1)

110
Z.G. Xu et al. Applied Thermal Engineering 137 (2018) 101–111

The functions I0 (z ),G0 (z ),I1 (z ),G1 (z ) are given by the following expressions:

z 2k
I0 (z ) = ∑
k=0
⎡ 12
⎣ (k !)
() 2


z 2k
G0 (z ) = −I0 (z )ln ( )+∑
z
2
k=0
⎡ψ (k + 1) 1 2
⎣ (k !) ()
2



z 2k − 1
I1 (z ) = ∑
k=0
⎡ 1
⎣ k ! (k + 1) !
()
2



z 2k + 1
G1 (z ) = I1 (z )ln ( )+
z
2
1 1

z 2

k=0
⎡ψ (k + 1) + ψ (k + 2) 1 2
⎣ (k !) ()
2

The function ψ (z ) is given by the following expression:


n ∞

ψ (z ) = limn →∞ ⎜ ∑ m−logn + ∑ ⎛ 1 − 1 ⎞⎞

⎝m=1 n=0 ⎝ n + 1 n + z ⎠⎠

8. The dimensionless bulk fluid temperature for formula (43)


2π 1
∫0 ∫0 Uθf RdRdφ Ri Rm 1
θfb = 2π 1
=2 ∫0 Uθf RdR + 2 ∫R Uθf RdR + 2 ∫R Uθf RdR
∫0 ∫ 0
URdRdφ i m

References and fluid temperature differentials in porous media, Int. J. Heat Mass Transf. 42
(1999) 423–435.
[16] W.H. Hsieh, J.Y. Wu, W.H. Shih, W.C. Chiu, Experimental investigation of heat
[1] Z.G. Xu, Z.G. Qu, C.Y. Zhao, Pool boiling heat transfer on open-celled metallic foam transfer characteristics of aluminum-foam heat sinks, Int. J. Heat Mass Transf. 47
sintered surface under saturation condition, Int. J. Heat Mass Transf. 54 (2011) (2004) 5149–5157.
3856–3867. [17] M. Nazari, M. Ashouri, M. Kayhani, A. Tamayol, Experimental study of convective
[2] Z.G. Xu, Z.G. Qu, C.Y. Zhao, Experimental study of pool boiling heat transfer on heat transfer of a nanofluid through a pipe filled with metal foam, Int. J. Therm. Sci.
metallic foam surface with U-shaped and V-shaped grooves, J. Enhanced Heat 88 (2015) 33–39.
Transf. 41 (2012) 44–55. [18] A. Hamadouche, R. Nebbali, H. Benahmed, A. Kouidri, A. Bousri, Experimental
[3] Z.G. Xu, C.Y. Zhao, Experimental study on pool boiling heat transfer in gradient investigation of convective heat transfer in an open-cell aluminum foams, Exp.
metal foams, Int. J. Heat Mass Transf. 85 (2015) 824–829. Therm. Fluid Sci. 71 (2016) 86–94.
[4] Z.G. Xu, C.Y. Zhao, Enhanced boiling heat transfer by gradient porous metals in [19] X. Xia, X. Chen, C. Sun, Z. Li, B. Liu, Experiment on the convective heat transfer
saturated pure water and surfactant solutions, Appl. Therm. Eng. 100 (2016) 68–77. from airflow to skeleton in open-cell porous foams, Int. J. Heat Mass Transf. 106
[5] Z.F. Huang, A. Nakayama, K. Yang, C. Yang, W. Liu, Enhancing heat transfer in the (2017) 83–90.
core flow by using porous medium insert in a tube, Int. J. Heat Mass Transf. 53 [20] M. Ghafarian, D. Mohebbi-Kalhori, J. Sadegi, Analysis of heat transfer in oscillating
(2010) 1164–1174. flow through a channel filled with metal foam using computational fluid dynamics,
[6] Y. Ge, Z. Liu, W. Liu, Multi-objective genetic optimization of the heat transfer for Int. J. Therm. Sci. 66 (2013) 42–50.
tube inserted with porous media, Int. J. Heat Mass Transf. 101 (2016) 981–987. [21] C. Chen, P. Huang, H. Hwang, Enhanced forced convective cooling of heat sources
[7] V.V. Calmidi, R.L. Mahajan, Forced convection in high porosity metal foams, ASME by metal-foam porous layers, Int. J. Heat Mass Transf. 58 (2013) 356–373.
J. Heat Transf. 44 (2001) 827–836. [22] H. Xu, L. Gong, S. Huang, M. Xu, Flow and heat transfer characteristics of nanofluid
[8] W. Lu, C.Y. Zhao, S.A. Tassou, Thermal analysis on metal-foam filled heat ex- flowing through metal foams, Int. J. Heat Mass Transf. 83 (2015) 399–407.
changers. Part I: metal-foam filled pipes, Int. J. Heat Mass Transf. 49 (2006) [23] H.J. Xu, Z.G. Qu, W.Q. Tao, Analytical solution of forced convective heat transfer in
2751–2761. tubes partially filled with metallic foam using the two-equation model, Int. J. Heat
[9] C.Y. Zhao, W. Lu, S.A. Tassou, Thermal analysis on metal-foam filled heat ex- Mass Transf. 54 (2011) 3846–3855.
changers. Part II: tube heat exchangers, Int. J. Heat Mass Transf. 49 (2006) [24] W. Lu, T. Zhang, M. Yang, Analytical solution of forced convective heat transfer in
2762–2770. parallel-plate channel partially filled with metallic foams, Int. J. Heat Mass Transf.
[10] Y.P. Du, Z.G. Qu, C.Y. Zhao, W.Q. Tao, Numerical study of conjugated heat transfer 100 (2016) 718–727.
in metal foam filled double-pipe, Int. J. Heat Mass Transf. 53 (2010) 4899–4907. [25] H.J. Xu, C.Y. Zhao, Z.G. Xu, Analytical considerations of slip flow and heat transfer
[11] S. Mahjoob, K. Vafai, Analysis of bioheat transport through a dual layer biological through microfoams in mini/microchannels with asymmetric wall heat fluxes, Appl.
media, ASME J. Heat Transf. 132 (2010) 031101.1-031101.14. Therm. Eng. 93 (2016) 15–26.
[12] K. Yang, K. Vafai, Analysis of temperature gradient bifurcation in porous media – an [26] C.Y. Zhao, T. Kim, T.J. Lu, H.P. Hodson, Thermal Transport Phenomena in Porvair
exact solution, Int. J. Heat Mass Transf. 53 (2010) 4316–4325. Metal Foams and Sintered Beds, Ph.D. report University of Cambridge, 2001.
[13] D. Poulikakos, M. Kazmierczak, Forced convection in duct partially filled with a [27] V.V. Calmidi, Transport Phenomena in High Porosity Fibrous Metal Foams, Ph.D.
porous material, ASME J. Heat Transf. 109 (1987) 653–662. Thesis University of Colorado, 1998.
[14] S. Chikh, A. Boumedien, K. Bouhadef, G. Lauriat, Analytical solution of non-Darcian [28] K. Boomsma, D. Poulikakos, On the effective thermal conductivity of a three-di-
forced convection in an annular duct partially filled with a porousmedium, Int. J. mensionally structured fluid-saturated metal foam, Int. J. Heat Mass Transf. 44
Heat Mass Transf. 38 (1995) 1543–1551. (2001) 827–836.
[15] D.Y. Lee, K. Vafai, Analytical characterization and conceptual assessment of solid

111

You might also like