You are on page 1of 12

Available online at www.sciencedirect.

com

Solar Energy 85 (2011) 2374–2385


www.elsevier.com/locate/solener

Coupled radiation and flow modeling in ceramic foam volumetric


solar air receivers
Zhiyong Wu a,⇑, Cyril Caliot b, Gilles Flamant b, Zhifeng Wang a
a
The Key Laboratory of Solar Thermal Energy and Photovoltaic System, IEE-CAS, Beijing 100190, China
b
Processes, Materials, and Solar Energy Laboratory, PROMES CNRS, 7 Rue du Four Solaire, 66120 Font-Romeu, France

Received 24 December 2010; received in revised form 23 June 2011; accepted 28 June 2011
Available online 23 July 2011

Communicated by: Associate Editor Avi Kribus

Abstract

Ceramic foams are promising materials for the absorber of volumetric solar air receivers in concentrated solar thermal power (CSP)
receivers. The macroscopic temperature distribution in the volumetric solar air receiver is crucial to guarantee that volumetric solar air
receivers work steadily, safely and above all, efficiently. This study analyzes the temperature distribution of the fluid and solid phases in
volumetric solar air receivers. The pressure drop in the ceramic foams and the interfacial heat transfer between the flowing fluid and solid
are included in the model. The radiative heat transfers due to concentrated solar radiation absorption by the ceramic foam and the radi-
ation transport in the media were modeled with the P1 approximation. The energy fields of the fluid and solid phases were obtained using
the local thermal non-equilibrium model (LTNE). Comparison of the macroscopic model with experimental results shows that the mac-
roscopic model can be used to predict the performance of solar air receivers. Sensitivity studies were conducted to analyze the effects of
velocity, porosity, mean cell size and the thermal conductivity of the solid phase on the temperature fields. The results illustrate that the
thermal non-equilibrium phenomena are locally important, and the mean cell size has a dominant effect on the temperature field.
Crown Copyright Ó 2011 Published by Elsevier Ltd. All rights reserved.

Keywords: Volumetric solar air receiver; Ceramic foam; P1 model; Local thermal non-equilibrium model; CFD

1. Introduction knowledge is linked with the solar air receivers0 safety, effi-
ciency, operability and the robustness of the whole solar
The large specific surface area of porous media and the tower power plant.
tortuous flow path inside the porous media make it a useful Studies of the temperature field inside porous media have
material for many industrial applications, especially in widely used the local thermal non-equilibrium model
applications where the heat transfer is important. Cellular (Alazmi and Vafai, 2002; Kim and Jang, 2002; Khashan
ceramics are promising materials for the absorbers of volu- et al., 2006; Nouri-Borujerdi et al., 2007; Hayes et al.,
metric solar receivers in concentrated solar thermal power. 2008). However, the radiation heat transfer has not been
The knowledge of the temperature difference between the taken into account in these studies. The radiation heat
flowing fluid and the solid matrix, and its influencing factors transfer plays a dominant role in the heat transfer when
are crucial to the entire system where the porous media is the porous is in a high temperature environment (Zhao
used as the heat exchanger. In CSP technology, this et al.,2004a,b). The working temperature of volumetric solar
receiver is very high, generally range from ambient tempera-
⇑ Corresponding author. Address: No. 6 Beiertiao, Zhongguancun, ture to 1000 °C or higher. Thus the radiation heat transfer is
100190 Beijing, China. Tel.: +86 10 82547037; fax: +86 10 62587946. important. Volumetric solar receivers have been studied for
E-mail address: wuzhiyongspecial@gmail.com (Z. Wu). more than 20 years. Flamant et al. (Flamant et al., 1988) and

0038-092X/$ - see front matter Crown Copyright Ó 2011 Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.solener.2011.06.030
Z. Wu et al. / Solar Energy 85 (2011) 2374–2385 2375

Nomenclature

C1,C2 k - e model constants Greek symbols


cp thermal capacity (J kg1 K1) a absorptivity
d mean cell size (m) r Stefan–Boltzmann constant
DNI direct normal irradiance (kW m2) b extinction coefficient (m1)
F source term representing the pressure drop in k thermal conductivity (Wm1 K1)
the porous media rs scattering coefficient (m1)
Gi generation rate of the intrinsic average of kt / porosity
hlv local volumetric heat transfer coefficient q density (kg m3)
(Wm3 K1) l dynamic viscosity (kg m1 s1)
I radiation intensity (Wm2 sr1) m kinematic viscosity (m2 s1)
K1 permeability coefficient (m2) e dissipation rate (m2 s3)
K2 inertial coefficient (m1) ee apparent emissivity
k absorption coefficient (m1) rk, re k  e model constants
kt turbulence kinetic energy (m2 s2)
Nulv Nusselt number based on hlv Subscripts
P pressure (Pa) c conduction
Pi production rate of hktii due to gradients of uD eff effective
q heat source (Wm3) f fluid
Q heat flux (Wm2) in inlet
Re Reynolds number (qud/l) out outlet
T temperature (K) r radiation

uD superficial velocity (m s1) s solid
u 0
velocity fluctuation (m s1) v volume
x x-direction coordinate (m) lv local and volumetric
z z-direction coordinate (m)
r cylindrical coordinate (m) Superscripts
i volume average

Variot et al. (Variaot et al., 1994) investigated the combined and thermal capacity) of ceramic foam materials and the
heat transfer in a two-slab selective volumetric solar air fluid properties jointly affect the performance of a volumet-
receiver made of a multilayer packed bed. Pitz-Paal et al. ric solar receiver. To design a volumetric solar air receiver
(Pitz-Paal et al., 1991) numerically investigated the air and with a favorable temperature distribution, we need to know
wall temperature distribution in a selective solar receiver the effects of these parameters on the volumetric solar recei-
which consisted of a ceramic foil receiver covered by a matrix ver performances. This study simulates the temperature dis-
of square channels of quartz glass. These volumetric solar tribution of the fluid and solid phases in a volumetric solar
receivers used the same selective absorptive concept in which air receiver by solving the coupled volume-averaged govern-
the maximum temperature locates inside of the volumetric ing equations. The studied ceramic foam, see Fig. 1, was
solar receiver. However, because of the complexity of the assumed isotropic, homogenous and with temperature
structures, the selective-absorbing volumetric solar receiver independent properties. The governing equations were
has not been studied much more. Researchers then switched volume-averaged. The pressure drop of ceramic foams,
to a volumetric solar receiver structure without a semitrans- the interfacial heat transfer between the flowing fluid and
parent layer. Fend et al. (Fend et al., 2004) proposed an ideal the solid, and the radiation heat transfer due the concen-
temperature distribution in the solid phase where the maxi- trated solar irradiance and inside the ceramic foams were
mum temperature is located inside the absorber. However, included. The pressure drop was calculated using a non-
experimental data in the literature (Fend et al., 2004) illus- Darcian model. The energy equations of the fluid and solid
trated that the maximum temperature is still located at the phases used a local thermal non-equilibrium model with the
front surface of the absorber. So this ideal temperature dis- concentrated solar irradiance been treated as a distributed
tribution is very difficult to attain. heat source to the solid phase. The thermal radiation heat
The geometric properties (mean cell size, porosity and transfer between strut surfaces was computed with the P1
the shape of the strut cross-section), optical properties model. The macroscopic model was then used to study
(absorptivity, extinction coefficient and scattering phase the impact of various factors (porosity, mean cell size,
function), thermo-physical properties (thermal conductivity etc.) on the temperature field. The macroscopic temperature
2376 Z. Wu et al. / Solar Energy 85 (2011) 2374–2385

Fig. 1. The studied ceramic foam material.

distributions of the fluid and solid phases in the volumetric negligible; and (e) because the absorber is far larger than
solar air receiver were obtained through this study. the mean cell size of the ceramic foam, the border effects
were neglected. With these assumptions, the volume-
2. Mathematical model averaged continuity and momentum equations are:
r  qf uD ¼ 0 ð1Þ
2.1. Flow modeling   
uD uD  
r  qf ¼ r /hP ii þ lr2 uD þ r
Transport phenomena in porous media are strongly /
affected by the complexity of the internal morphology of 
 /qf hu0 u0 ii þ /F ð2Þ
porous media, which is extremely complicated in general.
In addition, the pore size is usually 0.5–3 mm. These realities Where symbols hi mean the volumetric average.  uD is the
make quantitative studies of the flow and heat transfer in volume average velocity, which is also called the superficial
porous media very difficult, either by experiments or simula- velocity. qf u0 u0 is the Reynolds stress. F is the pressure drop

P
tions. However, engineers are usually only interested at the 3
resulting from porous media with F ¼  uD þ
j¼1 Dij l
macroscopic phenomena, instead of pore-scale phenomena P3 1
in porous media. Consequently, macroscopic descriptions uD juD Þ, where Dij and Cij are second-rank ten-
j¼1 C ij 2 qj
such as the volume averaging methodology, are the tool of sors. For homogenous porous media, these
two second-rank
choice (Lage et al., 2002). Indeed, with the macroscopic l
tensors are unit tensors, thus F ¼  K1
uD þ Kq2 j
uD j
uD Þ
model, all the complex geometry of porous media need not
be described. This approach greatly reduces the complexity where K1 is the permeability coefficient, and K2 is the iner-
of the physical problem and the cost of computations. The tial coefficient.
overall effects of the solid phase are taken into account in The macroscopic turbulence equations are also obtained
the macroscopic model by adding source terms to the by taking the volume average of the standard turbulence
governing equations of clear fluid. When considering the vol- kinetic energy and the dissipation rate equations:
ume-average (Quintard andWhitaker, 1994) of the governing   
 i l  i i
equations of the turbulent flow in porous media, there are r  qf uD hk t i ¼ r  l þ te r /hk t i þ P i þ Gi  qf /hei
rk
four available methodologies for developing macroscopic
turbulence models (Lage et al., 2002; De Lemos, 2005). How- ð3Þ
ever, some of the final equations are not yet closed or they   
 i lt e  i heii
can not be validated due to the lack of experimental data. r  qf uD hei ¼ r  l þ r /hei þ C 1 P i i
re hk t i
For this reason, the existing semi-empirical governing equa-
i
tions were used to describe the turbulent flow in porous hei  i
þ C2 i Gi  qf /hei ð4Þ
media (Pedras and de Lemos, 2001a,b; de Lemos, 2006; De hk t i
Lemos and Saito, 2008).
The following assumptions are made to address the where C1, C2, rk and re are k  e model constants, Pi is the
governing equations: (a) the porosity is uniform within production rate of hktii due to gradients of uD and Gi is the
the porous matrix; (b) the flow field is steady and turbulent, generation rate of the intrinsic average of kt due to the
and the absolute mean fluid velocity is in the range of action of the porous matrix (De Lemos and Saito,
1–15 m/s; (c) the thermo-physical properties of the solid 2008a,b).
  
phase are constant and the thermo-physical properties of   l  
the fluid phase are variable because of the large tempera- r  qf uD hk t ii ¼ r  l þ te r /hk t ii þ P i þ Gi
rk
ture range; (d) the effects of buoyancy, hydrodynamic i
dispersion, viscous dissipation and thermal expansion are  qf /hei ð5Þ
Z. Wu et al. / Solar Energy 85 (2011) 2374–2385 2377

   i
 i lte  i hei is considered as an absorbing, emitting and isotropic scatter-
r  qf 
uD hei ¼ r  lþ r /hei þ C 1 P i i ing media with homogeneous and gray radiative properties.
re hk t i
Furthermore, since the ceramic foam strongly absorbs the
heii  i solar radiation, the ceramic foam is considered to be opti-
þ C2 i Gi  qf /hei ð6Þ
hk t i cally thick with a short radiation transport mean free path.
Because of the high working temperature of the volu- The P1 model is suitable for optically thick media and is
metric solar receiver, the air properties are defined as vari- not time consuming. Therefore, the P1 model was used in this
ables in this study, namely the thermal capacity: study.
With a large optical thickness, the mean free path of
cp ¼ 1:93  1010 T 4f  8  107 T 3f þ 1:14  103 T 2f photons is very small compared to the mean beam length.
 4:49  101 T f þ 1:06  103 ð7Þ Consequently, the radiation intensity is only affected by the
radiation heat transfer in the near region and the integrated
The thermal conductivity: intensity tends to be isotropic. The integrated intensity is
then the solution of a diffusion equation. In this study,
k ¼ 1:52  1011 T 3f  4:86  108 T 2f þ 1:02  104 T f
the ceramic foam was irradiated by concentrated solar
 3:93  103 ð8Þ energy which was assumed to be a collimated incident radi-
ation beam. Thus, the P1 model was used with the intensity
Eqs. (5) and (6) are calculated from polynomial curve
splitting technique which uses collimated and diffuse inten-
fits to a data set for 100–1600 K in book ( Mulholland,
sities (Modest, 2003). As a result, the integrated intensity
1995). In addition, the viscosity was calculated through
computed by the P1 method was the diffuse irradiation
Sutherland Law. The air was treated as ideal gas when
Gs, i.e. the solar radiation that has undergone at least
computing the density and the gravity is neglected.
one scattering and the emitted radiation. The transport
equation of the diffuse integrated intensity Gs is,
2.2. Heat transfer modeling
 
1  
The governing equation for the energy conservation is a r  rGs ¼ k 4rT 4s  Gs þ rs I 0
3ðk þ rs Þ
volume-averaged equation. Two approaches, the local
thermal equilibrium model (LTE) and the local thermal  exp½bz ð11Þ
non-equilibrium model (LTNE), were widely used to treat
The collimated irradiation was computed using
the energy conservation in porous media. In this study, we
Gd = I0 exp[bz]. The total irradiation, G = Gs + Gd, was
used the LTNE model because this model can provide
then used to compute the radiative source term in Eq.
more temperature information of the air receiver. The
(9), r  qr ¼ k 4rT 4s  G .
LTNE has two energy equations for the fluid and solid
phases respectively:

3. Numerical model
i  i  i i
r  ðqcp Þf 
uD hT f i ¼ r  keff ;f  rhT f i þ hv hT s i  hT f i
3.1. Physical model
ð9Þ
 i  i i The sectional shape of the absorber in a real volumetric
0 ¼ r  keff ;s  rhT s i þ hv hT s i  hT f i þ r  qr ð10Þ
solar air receiver can be a rectangle, a hexagon or other
The energy coupling of the solid and fluid phases is real- shapes. Experimental studies have generally used absorber
ized by the volumetric convection term, hv, in their energy samples that were cylinders. In numerical studies, the
balance equations. A radiative source term is included in absorber shapes can be easily changed. For comparison
the energy balance equation for the solid phase, $  qr, with experimental data, one bulk of cylindrical shaped
which is the divergence of the radiative flux. ceramic foam was used in this study. The size of the foam
material, the physical model and the coordinate system are
2.3. Radiative transfer modeling shown in Fig. 2.

The difficulty of solving the radiative transfer equation 3.2. Effective conductivity
lies in the fact that the radiation intensity depends on the
position and on the direction. Hence, the deterministic meth- The thermal conductivity used in porous media is gener-
ods ( Modest, 2003) such as the Discrete Ordinates or the ally named as the effective thermal conductivity. The effec-
Finite Volume methods require computing the radiation tive conductivity is generally second-rank tensor. Owe to
intensity at each node of the domain and discretizing both the lack of specific knowledge about the exact effective
the domain and the direction of radiation propagation. conductivity and also for convenience, porous media mate-
The computational time associated with such methods can rials are usually treated as isotropic media, and the effective
become prohibitive, especially when strongly scattering conductivity is calculated through correlations. The correla-
media are involved. In the present study, the ceramic foam tion is generally a mixture of the conductivities of the clear
2378 Z. Wu et al. / Solar Energy 85 (2011) 2374–2385

Fig. 2. Schematic of the physical model.

fluid or the dense solid. The Schuetz–Glicksman formula Thus the extinction coefficient can be computed as:
(Schuetz and Glicksman, 1984), which Kamiuto considered
3ð1  /Þ
the most accurate among the formula examined in b ¼ k þ rs ¼ ð17Þ
d
( Kamiuto, 2008), is used in this study to determine the
effective thermal conductivity (note: the default method of
calculating the effective conductivity in FLUENT is 3.4. Pressure drop correlation
keff ¼ /kf þð1  /Þks ). The formulas are:
keff ;f ¼ /kf ð12Þ In the momentum equation in Eq (2), the term F is the
pressure drop resulting from the porous media. For homog-
1 enous porous media, this term involves the two parameters,
keff ;s ¼ ð1  /Þks ð13Þ
3 the permeability coefficient, K1, and the inertial coefficient,
1 K2. Many researchers have developed models (Richardson
keff ¼ /kf þ ð1  /Þks ð14Þ
3 et al., 2000; Fourie and Du Plessis, 2002a,b; Macdonald
et al., 2002; Lacroix et al., 2007; Incera Garrido et al.,
3.3. Optical properties of ceramic foams 2008; Wu et al., 2009, 2010) to correlate the two parameters
with the characteristic length and the porosity. Wu et al’s
When the volumetric absorber is at high temperatures, (Wu et al., 2009, 2010) model, which was developed using
the heat transport by radiation plays a significant role numerical simulation on idealized tetrakaidecahedral struc-
[25]. In addition, the concentrated solar incoming flux is tures and validated by experimental data, was used in this
mainly absorbed by the ceramic foam with typical working study. The relationships to describe the two parameters
temperatures of volumetric solar air receivers ranging from are K1 = d2/(1039  1002/) and K2 = 0.5138/5.739/d.
300 to 1300 K, thus the radiative transport is important.
The porous media studied in this work was assumed to 3.5. Interfacial heat transfer coefficient
be a homogeneous participating media whereas the air is
transparent. The emission, absorption and scattering by The volume-averaged governing equations are conve-
the porous media were considered. The radiative properties nient for engineering applications using porous media. The
needed for the modeling of the radiative transport in a par- LTNE model is used here to account for the temperature dif-
ticipating media are the extinction coefficient, albedo and ference between the fluid and solid phases. One difficulty of
scattering phase function. Like the effective thermal con- using the LTNE model is to accurately calculate the interfa-
ductivity in porous media, the optical properties have been cial heat transfer between the fluid and solid phases. Younis
extensively studied (Hale and Bohn, 1993; Hsu and Howell, (Younis and Viskanta, 1993), Fu (Fu et al., 1998), Hwang
1992; Mital et al., 1996; Baillis et al.,1999, 2000; Baillis and (Hwang et al., 2002) and Fend (Fend et al., 2005) used the
Sacadura, 2000; Tseng and Kuo, 2002; Petrasch et al., 2007; single-blow method to measure the interfacial heat transfer
Loretz et al., 2008; Bellet et al., 2009). In this study, the coefficient. Numerical simulations based on idealized geom-
absorption coefficient and the scattering coefficient were etries, such as the cubic model (Ghosh, 2008), a square cylin-
computed by assuming that the geometrical optics approx- der field (Kuwahara et al., 2006) and solid square rods (Saito
imation holds (Vafai, 2005): and de Lemos, 2005; Saito and Lemos, 2006), and real geom-
etries from computer tomography (Petrasch et al., 2008),
ð1  /Þ
k ¼ 1:5a ð15Þ have been used to model the flow and heat transfer. Wu
d et al. (Wu et al., 2009, 2010) numerically investigated the
ð1  /Þ convective heat transfer characteristics of air flow through
rs ¼ 1:5ð2  aÞ ð16Þ
d ceramic foam, and argued that it is not appropriate to use
Z. Wu et al. / Solar Energy 85 (2011) 2374–2385 2379

the “mean volumetric heat transfer coefficient”. They temperature (300 K), and Tz is the absorber’s front surface
instead proposed using the local volumetric heat transfer temperature.
coefficient given by,
3.7. Numerical method
Nulv ¼ ð32:504/0:38  109:94/1:38 þ 166:65/2:38
 86:98/3:38 ÞRe0:438 ð18Þ Although the physical model is a simple pipe, and cylin-
drical symmetric, a two dimensional mesh was generated
This model is for the local volumetric heat transfer coef-
and used in this study, see Fig. 3. This two dimensional
ficient in the fully developed flow region. In the entrance
conjugate heat transfer problem was solved using the
region, the local volumetric heat transfer coefficient is
CFD software FLUENT 6.3. The Reynolds number based
higher than the value predicted by Eq. (16). The entrance
on the mean cell size and the superficial velocity ranged
length is a function of the geometry, permeability, porosity
from 120 to 800 in the volumetric solar air receiver (Wu
and fluid viscosity. However, this length is generally only
et al., 2011). According to Kaviany (Kaviany, 1995), this
1–2 times the mean cell size (Wu et al., 2009, 2010); thus,
flow regime is usually turbulent and sometimes laminar.
the influence of the entrance length was neglected.
Kuwahara et al. (Kuwahara et al., 1998) investigated the
microscopic turbulence fields in a fluid-saturated periodic
3.6. Boundary conditions array using the low and high-Reynolds number versions
of the k  e models, and concluded that the differences in
3.6.1. Concentrated solar radiation the volume-averaged quantities predicted by both models
The incoming concentrated solar radiative flux was trea- are insignificant. In addition, the k  e model is so far
ted as parallel, constant and uniform when conducting the the only turbulence model being used in the volume-
sensitivity studies. However, the concentrated solar radia- average methodology. Thus, the standard k  e model
tive flux is actually not parallel, variable and non-uniformly (high Reynolds number) was used along with conventional
distributed. The numerical results were compared with wall functions with the LTNE models in this study. The
experimental data assuming a steady, parallel solar radia- model constants were as recommended by Launder and
tive flux with a Gaussian distribution. This Gaussian distri- Spalding, which are the default values in FLUENT.
bution was fit based on the measured data on the 6 kW The SIMPLE algorithm was used to couple the pressure
furnace in PROMES (Odeillo), CNRS, France. and the velocities with the second order upwind method
used for the advection terms in the momentum equations
3.6.2. Wall boundary conditions and for the energy equation. For the LTNE model, the
The walls were considered static, adiabatic, and diffuse additional scalar equations were solved through a user
gray surfaces. For the P1 model, the wall boundary condi- defined function (UDF). Sources terms were computed
tion was a Marshak’s boundary condition with a wall emis- through the UDF to account for the pressure drop and
sivity of unity (Ishimaru, 1999). to couple the energy equations of the fluid phase and solid
phases. The convergence criterion was 105 for all the
3.6.3. Inlet and outlet conditions equations.
The inlet boundary (see Fig. 2) was defined to have a
given pressure (ranging from 200 to 2000 Pa above the ref- 4. Results and discussion
erence pressure), and a static temperature of 300 K. The
outlet boundary was defined with a static pressure of 4.1. Sensitivity studies
0 Pa. When comparing with the experimental data, the ref-
erence pressure was set to 84,500 Pa and to 101,325 Pa For optimizing the performances of the volumetric solar
otherwise because the altitude of the experimental site is air receiver, we conducted a series of sensitivity study of the
about 1800 m. inlet velocity (by adjusting the inlet pressure), the porosity,
For the radiation, the radiation rays can penetrate into the mean cell size and the thermal conductivity. The sensi-
the porous media, and the emission from the strut surface tivity studies are based on the LTNE model and with the P1
inside the porous media can also go out. In addition, radiation model. The concentrated solar flux is a constant
because we split the concentrated solar irradiance into col- incident flux of 600 kW/m2, and the air inlet temperature
limated and diffuse intensities, the conventional Marshak’s of 300 K applies to all cases of this section.
boundary condition overestimates the radiation loss, and Fig. 4 presents the temperature distribution of the fluid
thus is not suitable for this study. Therefore, the boundary and solid phase with same absorber properties
condition for the radiative losses (Eq. (7)) was expressed as, (d = 1.5 mm, / = 0.8), but under different inlet superficial
 
qloss ¼ ee Sr T 4z  T 40 ð19Þ velocity. This figure shows that with the increasing of the
inlet velocity, the outlet temperature decreases and the
Where ee is the apparent emissivity of the ceramic foam whole solid temperature becomes more uniform. In addi-
(ee = 0.95, measured at ambient temperature), S the surface tion, the maximum temperature place locates always inside
area, r is the Stefan–Boltzmann constant, T0 is the ambient the absorber.
2380 Z. Wu et al. / Solar Energy 85 (2011) 2374–2385

Air in Air out

Axis

Fig. 3. Mesh used in this study.

Fig. 4. Solid and fluid temperature distributions for various inlet velocities Fig. 5. Solid and fluid temperature distributions for various material
(under the same geometrical properties of the material and solar porosity (under the same inlet mass flow rate and solar irradiance).
irradiance).

Fig. 6 presents the temperature distribution of the fluid


Fig. 5 presents the temperature distribution of the fluid and solid phase with different mean cell size (/ = 0.8). A
and solid phase under different porosity (d = 1.5 mm). A fixed mass flow rate was defined as the inlet boundary con-
fixed mass flow rate was defined as the inlet boundary dition. This figure shows that the mean cell size has also a
conditions. This figure shows that the porosity has a remarkable influence on the front temperature of the
remarkable influence on the front temperature of the absorber. Like the influences of the porosity, the most
absorber. The maximum temperature difference on obvious difference is the front temperature of the receiver.
the front temperature of the receiver is about 150 K when The temperature difference on the front temperature of the
the porosity varies from 0.7 to 0.9. The reason is that the receiver is about 50 K when the mean cell size varies from
porosity has a considerable influence on the extinction 1.0 mm to 3.0 mm. The reason is the extinction coefficient
coefficient of foam materials. The extinction coefficient decreases reversely with the mean cell size. Therefore, the
decreases reversely with the porosity. Therefore, the con- concentrated solar irradiance can penetrate into the absor-
centrated solar irradiance can penetrate into the absorber ber to a more depth when the cell size increases.
to a more depth when the porosity increases. In addition, To compare Figs. 5 and 6, we can see that the porosity
the porosity has a weak influence on the interfacial heat has a higher influence on the front temperature of the
transfer coefficient. As a result, the front temperature of absorber, whereas the cell size has a higher influence on
the absorber decreases reversely with the porosity. For the thickness of the thermal non-equilibrium region. In
the same reason, the thickness of the thermal non- addition, the difference of the fluid temperature distribu-
equilibrium region increases slowly with the porosity. tion is clearer in Fig. 6. This is because the cell size has a
Z. Wu et al. / Solar Energy 85 (2011) 2374–2385 2381

conductivity. Above all, it is possible to design a volumetric


solar air receiver with the ideal temperature distribution
described in (Fend et al., 2004) and called the “volumetric
effect”.
One feasible way to optimize the air receiver is to mini-
mize the thermal radiation losses by decreasing the absor-
ber’s front temperature. For example, when the receiver
works at 1000 K, a temperature difference of 100 K at the
absorber’s front surface means a difference of 5% of the
receiver’s thermal efficiency. However, the overall efficiency
is another thing. Fig. 8 presents the relationship between
the pressure drop of the receiver and the cell size, and
porosity. This figure shows that higher porosity is good
for the receiver, however, smaller cell size will cause higher
pressure drop.
Fig. 6. Solid and fluid temperature distributions for various mean cell From the above analysis, we can conclude that a type of
sizes (under the same inlet mass flow rate and solar irradiance).
ceramic foam, with a mean cell size of about 1–2 mm and a
highest porosity, is the most favorable choice for the absor-
higher influence on the interfacial heat transfer coefficient ber micro-structure. Of course, the final decision of the
than that of the porosity. absorber micro-structure is dependent on many factors,
Fig. 7 presents the temperature distribution of the fluid such as the mechanical properties and the manufacture
and solid phase with different thermal conductivity of the techniques, instead of only the receiver’s efficiency.
solid phase, and under the same inlet mass flow rate and
solar irradiance. This figure shows that the influences on 4.2. Comparison with experiment
the temperature field from the solid thermal conductivity
is very small when the air receiver works under steady state. The above macroscopic model were compared using
The maximum temperature difference is only 5 K when the experimental data from the literature (Wu, 2010), in which
thermal conductivity varies from 40 to 100 W/(m K) (note: a laboratory volumetric open solar air receiver was tested
this conductivity is not the effective conductivity, it is the on a solar furnace. The absorber material of this receiver
thermal conductivity of the pure solid). Consequently, we was ceramic foam and made of SiC. The experimental
can say that the solid thermal conductivity has small influ- setup used to study the absorber is shown in Fig. 9. In this
ence on the performances of a volumetric solar air receiver experiment, an infrared camera was used to measure the
when it works under steady states. absorber’s inlet surface temperature, two thermocouples
From these sensitivity studies, we can see that the influ- were used to measure the absorber’s inside temperature
ences from the velocity, porosity, mean cell size and conduc- (z = 25 mm, one locates at radius = 0 mm, the other
tivity on the temperature fields of the air receiver are very locates at radius = 12.5 mm), and two thermocouples were
complicated. At steady operating conditions, the influences used to measure the absorber’s outlet surface temperature
from the geometrical properties (mean cell size, porosity) (z = 50 mm, one locates at radius = 0 mm, the other
are predominant, and we can ignore the effects from the solid locates at radius = 12.5 mm). One thing should be stressed

Fig. 7. Solid and fluid temperature distributions for various solid thermal Fig. 8. Relationship between the pressure drop and mean cell size,
conductivities (under the same inlet mass flow rate and solar irradiance). porosity (under the same inlet mass flow rate and solar irradiance).
2382 Z. Wu et al. / Solar Energy 85 (2011) 2374–2385

Concentrator
Compressor
Flowmeter Quartz
tube

valve

insolator

pressure

PC
Data acquisition
Support
structure
Infrared camera sample 5 thermocouples

Hot air emit to


environment directly

Sun

Heliostat

Fig. 9. Experimental system (Wu, 2010).

here is that a preliminary test was conducted, and the can be analyzed as steady. And the experimental data are
results verified the thermocouple indications are corre- comparable to the numerical results.
sponding to the air temperature, and also the solid temper- The comparison was made by comparing the measured
ature when the thermocouples were installed at any temperature with the numerical results. The locations were
locations of the sample only if z > 25 mm, because where selected at the front surface, and the center line of the
the fluid phase has reached local thermal equilibrium with absorber. The temperature distribution in the solid phase
the solid phase. along a diameter line on the front surface is shown in
Experimental data of one sample were selected to com- Fig. 10a with the temperature distribution in the solid
pare with the macroscopic model developed in this study. phase along the center line shown in Fig. 10b. We attribute
The absorber properties are: porosity 0.7, cell size the discrepancy mainly to the measuring uncertainty of the
2.2 mm, diameter 50 mm and thickness 50 mm. The corre- infrared camera, because we define the emissivity of the
sponding operating conditions are: mass flow rate 3.36 g/s, absorber as unity during the test. Anyway, the computed
inlet air temperature 295 K, pressure drop 1760 Pa. results agree well with the experimental data with errors
To compare with the experimental data, the boundary in the solid temperature of less than 10%. Thus, the numer-
conditions and the absorber properties of the numerical sim- ical simulation is very successful and the LTNE model
ulation were adjusted according to the operating conditions developed in this study is appropriate.
in the experimental studies. The concentrated solar radiation The predicted temperature distributions of the fluid and
was assumed to be parallel beams with a Gaussian distribu- solid phases are shown in Fig. 11. We can see that at the
2
tion Q ¼ DNI  5:4289e2:873ðr=0:0335Þ kW=m2 . Because of the outlet position, the fluid and solid phases have reached
reflection, absorption and shadowing due to the quartz tube local thermal equilibrium. This result agrees with the
and supporting structures (see Fig. 9), 37.8% of the radiation experimental measurements which show that the signals
is attenuated according to the calorimetric measurement. from the two thermocouples almost overlap each other
Thus we assume that the solar flux on the sample (Wu, 2010).
2
is Q ¼ DNI  ð1  0:378Þ  5:4289 e2:873ðr=0:0335Þ kW=m2 .
The experiments were transient because the solar irradiance 5. Conclusions and future work
continuously varied, but the authors averaged the original
data over 2–5 min to get quasi-steady results. In this way, The combined heat transfer in volumetric solar air
the operating conditions of the volumetric solar receiver receivers was numerically simulated. The local thermal
Z. Wu et al. / Solar Energy 85 (2011) 2374–2385 2383

(a) 1800

1600
CFD
EXP
1400

Temperature (K) 1200

1000

800

600

400
-0.025 -0.0125 0 0.0125 0.025
Position (m)

(b) 1800

1600
CFD
EXP
Temperature (K)

1400

1200

1000

800
0 0.01 0.02 0.03 0.04 0.05
Position (m)

Fig. 10. Measured and predicted temperatures of the solid phase (a) along a radial line on the front surface, and (b) alone the center line of the absorber.

non-equilibrium model was used with the P1 model to (4) The thermal non-equilibrium region in the absorber
investigate the temperature distributions. A sensitivity is rather thin for ceramic foam with mean cell sizes
study was conducted to study the effect of the inlet velocity, 1–3 mm.
porosity, mean cell size and thermal conductivity. The sim-
ulations were also compared with experimental data. The This macroscopic method and the developed macro-
results are summarized below: scopic model in this paper are very helpful to the study
of the flow and heat transfer in porous media and the
(1) A generalized macroscopic model has been developed design of volumetric solar air receivers. However, these
for the modeling of ceramic foam used in the volu- conclusions were drawn from steady state simulations
metric solar air receivers. and the experimental data are not enough to fully validate
(2) The ideal temperature distribution with the maximum the proposed model. In addition, because the concentrated
solid temperature located inside the absorber is realiz- solar radiative flux in the focal spot of the furnace is far lar-
able. This temperature distribution is dependent on the ger than the solar radiative flux used in the section of sen-
working conditions and the material properties. sitivity studies, the sensitivity analysis leads to lower fluid
(3) The thermal conductivity of the solid phase of the and solid temperatures than those observed in the experi-
absorber material is not important to the performance mental study. In the future, more experimental and numer-
of the volumetric solar air receiver at steady state. ical studies will be conducted to validate the macroscopic
2384 Z. Wu et al. / Solar Energy 85 (2011) 2374–2385

0.025 De Lemos, M.J.S., Saito, M.B., 2008b. Computation of turbulent heat


K
transfer in a moving porous bed using a macroscopic two-energy
equation model. International Communications in Heat and Mass
1596 0.0125 Transfer 35 (10), 1262–1266.
Fend, T., Pitz-Paal, R., et al., 2004. Two novel high-porosity materials as
1452 volumetric receivers for concentrated solar radiation. Solar Energy
0 Ts
Y

Materials and Solar Cells 84 (1–4), 291–304.


1308 Fend, T., Trimis, D., et al. (Eds.), 2005. Thermal Properties. Cellular
-0.0125 Ceramics: Structure, Manufacturing, Properties and Applications.
1164 Weinheim, WILEY-VCH Verlag GmbH & Co. KGaA.
Flamant, G., Menigault, T., et al., 1988. Combined heat transfer in a
-0.025 semitransparent multilayer packed bed. Journal of Heat Transfer 110
1020 0 0.01 0.02 0.03 0.04 0.05
0.025 X (2), 463–467.
Fourie, J.G., Du Plessis, J.P., 2002a. Pressure drop modelling in cellular
876 metallic foams. Chemical Engineering Science 57 (14), 2781–2789.
0.0125 Fourie, J.G., Du Plessis, J.P., 2002b. Pressure drop modelling in cellular
732 metallic foams. Chemical Engineering Science 57 (14), 2781–2789.
Fu, X., Viskanta, R., et al., 1998. Measurement and correlation of
0
Y

Tf volumetric heat transfer coefficients of cellular ceramics. Experimental


588
Thermal and Fluid Science 17 (4), 285–293.
Ghosh, I., 2008. Heat-transfer analysis of high porosity open-cell metal
444 0.0125 foam. Journal of Heat Transfer 130, 034501.
Hale, M.J., Bohn, M.S., 1993. Measurement of the radiative transport
300 properties of reticulated alumina foams. In: ASME/ASES Joint Solar
-0.025
0 0.01 0.02 0.03 0.04 0.05 Energy Conference, ASME, New York, pp. 507–515.
X Hayes, A.M., Khan, J.A., et al., 2008. The thermal modeling of a matrix
heat exchanger using a porous medium and the thermal non-
Fig. 11. Predicted temperature distribution in the absorber under the solar
equilibrium model. International Journal of Thermal Sciences 47
irradiation of a Gaussian distribution.
(10), 1306–1315.
Hsu, P., Howell, J.R., 1992. Measurements of thermal conductivity and
model within larger velocity and temperature ranges, and optical properties of porous partially stabilized zirconia. Experimental
Heat Transfer 5 (4), 293–313.
the macroscopic model will be extended to the unsteady
Hwang, J.J., Hwang, G.J., et al., 2002. Measurement of interstitial
state conditions. convective heat transfer and frictional drag for flow across metal
foams. Journal of Heat Transfer 124, 120–129.
Acknowledgment Incera Garrido, G., Patcas, F.C., et al., 2008. Mass transfer and pressure
drop in ceramic foams: a description for different pore sizes and
porosities. Chemical Engineering Science 63 (21), 5202–5217.
This project was financially supported by the National
Ishimaru, A., 1999. Wave Propagation and Scattering in Random Media.
Basic Research Program of China (2010CB227106). Wiley-IEEE Press.
Kamiuto, K., 2008. Modeling of composite heat transfer in open-cellular
References porous materials at high temperatures. In: Chsner, A., Murch, G., De
Lemos, M. (Eds.), Cellular and Porous Materials: Thermal Properties
Alazmi, B., Vafai, K., 2002. Constant wall heat flux boundary conditions Simulation and Prediction. Wiley-VCH.
in porous media under local thermal non-equilibrium conditions. Kaviany, M., 1995. Principles of Heat Transfer in Porous Media. Springer.
International Journal of Heat and Mass Transfer 45 (15), 3071–3087. Khashan, S.A., Al-Amiri, A.M., et al., 2006. Numerical simulation of
Baillis, D., Sacadura, J.F., 2000. Thermal radiation properties of dispersed natural convection heat transfer in a porous cavity heated from below
media: theoretical prediction and experimental characterization. Journal using a non-Darcian and thermal non-equilibrium model. Interna-
of Quantitative Spectroscopy and Radiative Transfer 67 (5), 327–363. tional Journal of Heat and Mass Transfer 49 (5–6), 1039–1049.
Baillis, D., Raynaud, M., et al., 1999. Spectral radiative properties of Kim, S.J., Jang, S.P., 2002. Effects of the Darcy number, the Prandtl
open-cell foam insulation. Journal of Thermophysics and Heat number, and the Reynolds number on local thermal non-equilibrium.
Transfer 13 (3), 292–298. International Journal of Heat and Mass Transfer 45 (19), 3885–
Baillis, D., Raynaud, M., et al., 2000. Determination of spectral radiative 3896.
properties of open cell foam: model validation. Journal of Thermo- Kuwahara, F., Kameyama, Y., et al., 1998. Numerical modeling of
physics and Heat Transfer 14 (2), 137–143. turbulent flow in porous media using a spatially periodic array.
Bellet, F., Chalopin, E., et al., 2009. RDFI determination of anisotropic Journal of Porous Media 1, 47–56.
and scattering dependent radiative conductivity tensors in porous Kuwahara, F., Yamane, T., et al., 2006. Large eddy simulation of
media: application to rod bundles. International Journal of Heat and turbulent flow in porous media. International Communications in
Mass Transfer 52 (5–6), 1544–1551. Heat and Mass Transfer 33 (4), 411–418.
De Lemos, M.J.S., 2005. The double-decomposition concept for turbulent Lacroix, M., Nguyen, P., et al., 2007. Pressure drop measurements and
transport in porous media. In: Ingham, D., Pop, I. (Eds.), Transport modeling on SiC foams. Chemical Engineering Science 62 (12),
Phenomena in Porous Media III. Pergamon. 3259–3267.
de Lemos, M., 2006. Turbulence in Porous Media: Modeling and Lage, J.L., De Lemos, M.J.S., et al., 2002. Modeling turbulence in porous
Applications. Elsevier Science Ltd. media. In: Ingham, D., Pop, I. (Eds.), Transport Phenomena in Porous
De Lemos, M., Saito, M., 2008a. Computation of turbulent heat transfer Media II. Pergamon.
in a moving porous bed using a macroscopic two-energy equation Loretz, M., Coquard, R., et al., 2008. Metallic foams: radiative properties/
model. International Communications in Heat and Mass Transfer 35 comparison between different models. Journal of Quantitative Spec-
(10), 1262–1266. troscopy and Radiative Transfer 109 (1), 16–27.
Z. Wu et al. / Solar Energy 85 (2011) 2374–2385 2385

Macdonald, I.F., El-Sayed, M.S., et al., 2002. Flow through porous Saito, M.B., de Lemos, M.J.S., 2005. Interfacial heat transfer coefficient
media-the Ergun equation revisited. Industrial and Engineering for non-equilibrium convective transport in porous media. Interna-
Chemistry Fundamentals 18 (3), 199–208. tional Communications in Heat and Mass Transfer 32 (5), 666–676.
Mital, R., Gore, J., et al., 1996. Measurements of radiative properties of Saito, M.B., Lemos, M.J.S.d., 2006. A correlation for interfacial heat
cellular ceramics at high temperatures. Journal of Thermophysics and transfer coefficient for turbulent flow over an array of square rods.
Heat Transfer 10 (1), 33–38. Journal of Heat Transfer 128, 444–452.
Modest, M., 2003. Radiative Heat Transfer. Academic Pr. Schuetz, M.A., Glicksman, L.R., 1984. A basic study of heat transfer
Mulholland, G.W., 1995. Smoke production and properties. SFPE through foam insulation. Journal of Cellular Plastics 20 (2), 114.
Handbook of Fire Protection Engineering. National Fire Protection Tseng, C., Kuo, K., 2002. Thermal radiative properties of phenolic foam
Associationss, Quincy, MA. insulation. Journal of Quantitative Spectroscopy and Radiative
Nouri-Borujerdi, A., Noghrehabadi, A.R., et al., 2007. The linear stability Transfer 72 (4), 349–359.
of a developing thermal front in a porous medium: the effect of local Vafai, K., 2005. Handbook of Porous Media. CRC.
thermal non-equilibrium. International Journal of Heat and Mass Variaot, B., Menigault, T., et al., 1994. Modelling and optimization of a two-
Transfer 50 (15–16), 3090–3099. slab selective volumetric solar receiver. Solar Energy 53 (4), 359–368.
Pedras, M.H.J., de Lemos, M.J.S., 2001a. Macroscopic turbulence model- Wu, Z., 2010. Studies of the Flow and Heat Transfer in Ceramic Foams in
ing for incompressible flow through undeformable porous media. the Application of Solar Air Receiver. PhD Thesis. The Graduate
International Journal of Heat and Mass Transfer 44 (6), 1081–1093. School of Chinese Academy of Sciences, Beijing.
Pedras, M.H.J., de Lemos, M.J.S., 2001b. Simulation of turbulent flow in Wu, Z., Caliot, C., et al., 2009. Numerical study of interfacial heat transfer
porous media using a spatially periodic array and a low Re two- between air flow and ceramic foams for optimizing volumetric solar air
equation closure. Numerical Heat Transfer. Part A, Applications 39 receiver. In: Proceedings of the International Solar PACES Sympo-
(1), 35–59. sium on Consentrating Solar Power and Chemical Energy Systems.
Petrasch, J., Wyss, P., et al., 2007. Tomography-based Monte Carlo Berlin, Germany.
determination of radiative properties of reticulate porous ceramics. Wu, Z., Caliot, C., et al., 2010. Experimental and numerical study on
Journal of Quantitative Spectroscopy and Radiative Transfer 105 (2), pressure drop in ceramic foams for volumetric solar receiver applica-
180–197. tions. Apllied Energy 87, 504–513.
Petrasch, J., Meier, F., et al., 2008. Tomography based determination of Wu, Z., Caliot, C., et al., 2011. Numerical simulation of convective heat
permeability, Dupuit–Forchheimer coefficient, and interfacial heat transfer between air flow and ceramic foams to optimise volumetric
transfer coefficient in reticulate porous ceramics. International Journal solar air receiver performances. International Journal of Heat and
of Heat and Fluid Flow 29 (1), 315–326. Mass Transfer 54 (7–8), 1527–1537.
Pitz-Paal, R., Morhenne, J., et al., 1991. A new concept of a selective solar Younis, L.B., Viskanta, R., 1993. Experimental determination of the
receiver for high temperature applications. Solar Energy Materials 24 volumetric heat transfer coefficient between stream of air and ceramic
(1–4), 293–306. foam. International Journal of Heat and Mass Transfer 36 (6), 1425–1434.
Quintard, M., Whitaker, S., 1994. Transport in ordered and disordered Zhao, C.Y., Lu, T.J., et al., 2004a. Thermal radiation in ultralight metal
porous media II: generalized volume averaging. Transport in Porous foams with open cells. International Journal of Heat and Mass
Media 14 (2), 179–206. Transfer 47 (14–16), 2927–2939.
Richardson, J.T., Peng, Y., et al., 2000. Properties of ceramic foam Zhao, C.Y., Lu, T.J., et al., 2004b. The temperature dependence of
catalyst supports: pressure drop. Applied Catalysis A, General 204 (1), effective thermal conductivity of open-celled steel alloy foams. Mate-
19–32. rials Science and Engineering A 367 (1–2), 123–131.

You might also like