You are on page 1of 15

Applied Energy 229 (2018) 700–714

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Comparison of direct numerical simulation with volume-averaged method T


on composite phase change materials for thermal energy storage

Xiaohu Yanga,b,c, Qingsong Baia, Zengxu Guoa, Zhaoyang Niua, Chun Yangd, Liwen Jina, ,

Tian Jian Lue, , Jinyue Yanb,c
a
Institute of the Building Environment & Sustainability Technology, School of Human Settlements and Civil Engineering, Xi’an Jiaotong University, Xi’an 710049, PR China
b
Department of Chemical Engineering and Technology/Energy Processes, Royal Institute of Technology (KTH), 100 44 Stockholm, Sweden
c
School of Sustainable Development of Society and Technology, Mälardalen University, 721 23 Västerås, Sweden
d
School of Mechanical and Aerospace Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore
e
State Key Laboratory of Mechanics and Control of Mechanical Structures, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, PR China

H I GH L IG H T S

• Volume-averaged method is compared with direct numerical simulation in phase change.


• DNS results reports significant temperature difference between ligaments and PCM.
• Natural convection strongly affects the melting front shape and temperature distribution.
• VAM with two-temperature model predicts well, although pore-scale features are lost.

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

Keywords: Melting heat transfer in open-cell metal foams embedded in phase-change materials (PCMS) predicted by the
Open-cell metal foam volume-averaged method (VAM) was systematically compared with that calculated using direct numerical si-
Volume averaged method mulation (DNS), with particular attention placed upon the contribution of natural convection in the melt region
Direct numerical simulation to overall phase change heat transfer. The two-temperature model based on the assumption of local thermal non-
Phase change
equilibrium was employed to account for the large difference of thermal conductivity between metallic liga-
Natural convection
ments and PCM (paraffin). The Forchheimer extended Darcy model was employed to describe the additional
flow resistance induced by metal foam. For the DNS, a geometric model of metal foam based on tetra-
kaidehedron cells was reconstructed. The DNS results demonstrated significant temperature difference between
ligament surface and PCM, thus confirming the feasibility of local thermal non-equilibrium employed in VAM
simulations. Relative to the DNS results, the VAM combined with the two-temperature model could satisfactorily
predict transient solid-liquid interface evolution and local temperature distribution, although pore-scale features
of phase change were lost. The presence of natural convection affected significantly the melting front shape,
temperature distribution and full melting. The contribution of natural convection to overall phase change heat
transfer should be qualitatively and quantitatively given sufficient consideration from both macroscopic (VAM)
and microscopic (DNS) point of views. Besides, practical significance and economic prospective using metal
foam in TES unit for WHR system to provide residential heating or hot water is discussed and analyzed.

1. Introduction by 30% while coal and natural gas will increase by 50% from 2007 to
2035. China is the biggest country for energy consumption worldwide
The energy demand has increased sharply in the world in recent with an increase rate of 18% in 2011, while the US, EU, India, Russia
decades according to a summary of energy consumption for 69 coun- and Brazil follow behind. Research demonstrates that the energy con-
tries, especially in developing countries that are accelerating urbani- sumption in China will increase to be more than 15 times by 2050
zation process [1]. It is indicated that the demand for oil will increase compared with that in 1970 [2]. The rapid consumption of


Corresponding authors.
E-mail addresses: lwjin@xjtu.edu.cn (L. Jin), tjlu@xjtu.edu.cn (T.J. Lu).

https://doi.org/10.1016/j.apenergy.2018.08.012
Received 9 October 2017; Received in revised form 28 July 2018; Accepted 3 August 2018
0306-2619/ © 2018 Elsevier Ltd. All rights reserved.
X. Yang et al. Applied Energy 229 (2018) 700–714

Nomenclature u Component velocity in x axis (m s−1)



U Velocity vector (m s−1)
Abbreviation v Component velocity in y axis (m s−1)
V Void volume for a TES unit (m3)
DNS Direct numerical simulation Vc Fully charged volume for a TES unit (m3)
Expt Experimental measurement w Component velocity in z axis (m s−1)
MF Metal foam 〈〉 Extrinsic average of a quantity over a control volume
NS Numerical simulation || Magnitude of a vector
TES Thermal energy storage
PCM Phase change material Greek symbols
PF Plate fin
PPI Pore per inch α Cross-sectional area ratio of node to solid ligament
VAM Volume-averaged method αsf Specific area (m−1)
WHR Waste heat recovery β Thermal expansion coefficient (K−1)
δ Numerical constant
Symbols ε Porosity
η Percentage of welding to materials cost
Am Numerical coefficient for damping velocity ρ Density (kg m−3)
CE Inertia coefficient (m−1) μ Dynamic viscosity (N s m−2)
Cp Specific heat (J kg−1 K−1) Π Yearly earned profit ($·yr−1)
Ch Price for residential heat ($·GJ−1) σ Liquid fraction in the porous medium
Cut Specific material cost for metal foam ($·m−3) χ Flow tortuosity
dp Pore diameter (m)
e Thickness ratio of node to solid ligament Subscript
fl Melting fraction
Fo Fourier number e Effective parameter
G Shape function for metallic ligaments f Phase change material
hsf Interstitial heat transfer coefficient (W m−2 K−1) full Full melting rate
k Thermal conductivity (W m−1 K−1) i Initial state
L Latent heat of fusion of PCM (J kg−1) m Melting point
N Number of operating hours (h) s Solid ligaments
Nu Nusselt number td Thermal dispersion
Nu Integral mean Nusselt number w Wall
P Pressure (Pa) 1 Solid state
Qc Total stored energy (J) 2 Liquid state
T Temperature (K)
t Time (s)

conventional fossil fuels has therefore caused a lot of serious energy and enhancement [10–13]. A series of studies have indicated that metal
environmental issues, such as energy shortages, global warming, en- foams possess a superior performance in enhancing the thermal con-
vironmental pollution and etc. [3–5]. For the time being, convectional ductivity of PCMs during phase change process. To model the phase
fossil fuels have still played the vital role in global energy supply. To change process in a PCM-foam composite, two approaches are com-
mitigate the environment stress, on one hand, efforts have been paid to monly employed according to the scale considered in the simulations:
transferring the fossil fuel energy sources towards renewable energy, the volume-averaged method (VAM) and the direct numerical simula-
such as solar, wind, tidal and biomass and so on; on the other hand, tion (DNS).
turning “waste heat” to “useful heat” can also improve the overall en- The VAM typically treats the PCM-foam composite as an equivalent
ergy efficiency of the existed energy systems, since about 20–50% of fluid with identical thermophysical properties, with pore-scale thermal
energy used in industry is rejected as waste heat [6,7]. fluid and geometric features lumped into the equivalent fluid. The
Waste heat recovery (WHR) system is capable of collecting in- predicted results demonstrate global fluid flow and heat transfer char-
dustrial “waste heat” and converting it to “useful heat” for incoming acteristics. Assuming local thermal equilibrium or non-equilibrium, one
reactants preheating, electricity generation, hot water and central may employ the one- or two-temperature model. Pioneering work on
heating for residences, and so on. Thermal energy storage (TES) via thermal equilibrium assumption (one-temperature model) can be traced
absorbing/releasing latent heat by phase change materials (PCMs) is back to Beckermann and Viskanta [14], who applied volume-averaged
one of the proved approaches for WHR. However, the charging/dis- transport equations to model the process of phase change in porous
charging rate suffers significantly from the low thermal conductivity of media. The topology of solid-liquid interface was found to be greatly
engineering-utilized PCMs (e.g. paraffin ∼ 0.2 Wm−1 K−1) [8], thus influenced by the onset of natural convection in the melt phase. Feng
limiting its practical applications. To this end, various techniques have et al. [15] and Yang et al. [16] applied the one-temperature model to
been developed to enhance phase change process. To be summarized, simulate the phase change behavior of PCM saturated in open-cell metal
there are two main approaches according to the thermal enhancement foam, and found that the model may be suitable for solidification
spreader adopted: movable – micro/nano particles, or non-movable – process.
porous matrix. Compared with the attainable enhancement by directly As for melting heat transfer in porous media, the two-temperature
adding micro/nano particles into the PCMs [9], the impregnation of model that considers the temperature difference between solid ligament
PCMs into highly porous metal foams with open cells is particularly surface and PCM may be more appropriate [17]. Assuming local
promising for structural controllability and satisfactory thermal thermal non-equilibrium, Mesalhy et al. [18] considered the Darcy-

701
X. Yang et al. Applied Energy 229 (2018) 700–714

Brinkman-Forchiemer effect on free convection in the melt region. The distribution and temperature difference between metallic ligament
phase change behavior was analyzed for different Fourier numbers and surface and PCM. However, thus far, pore-scale numerical simulations
the predictions agreed well with those in literature. Since heat transfer are scarce [29–34]. Notably, Hu et al. [29,32] conducted direct nu-
between ligament surface and the interstitial fluid was difficult to merical simulation of the melting heat transfer in open-cell metal foams
measure, the heat transfer coefficient across a bank of cylinders was having idealized cellular topologies: a body-centered-cubic (BCC) unit
commonly employed. By varying the interstitial heat transfer coeffi- cell was constructed to model the phase change process with constant
cient, Krishnan et al. [19] discussed the effect of interstitial Nusselt thermal boundary condition. With temperature distribution and phase
number on the propagation of melting front. Besides, the influence of interface evolution illustrated and the dominant role of heat conduction
Rayleigh and Stefan numbers on transient phase change heat transfer identified, they confirmed that the DNS approach acquired affordable
was also documented and discussed. Chen et al. [20] investigated computation resources and time. Sundarram and Li [30] and Yao et al.
thermal energy storage in a solar flat-plate collector with metal foam as [33] developed a 3D (three-dimensional) finite element (FE) model to
heat transfer enhancer for phase change. It was demonstrated that, in calculate the effective thermal conductivities of PCM-metal-foam
comparison with the one-temperature model, the two-temperature composites. The effects of pore size and porosity upon effective con-
model can better predict phase change heat transfer in the solar thermal ductivity as well as melting heat transfer characteristics were quanti-
collector. Tian and Zhao [21] solved the problem of coupled heat fied. Feng et al. [31] and Zhang et al. [34] also conducted pore-scale
conduction and natural convection in phase transition and liquid zones numerical simulations on phase change heat transfer in metal foams,
for PCM-metal-foam composites. They confirmed the applicability of and suggested that the melting heat transfer could be further enhanced
local thermal non-equilibrium model to simulate melting heat transfer with extra fin inserts and gradient packaging. Besides, PCM was also
in porous materials that have much higher thermal conductivity than impregnated into graphite foam to form graphite-foam-PCM composites
PCMs. Li et al. [22] conducted both experimental and numerical in- [35–37]. With graphite foam involved, the effective thermal con-
vestigations on melting phase change in copper foams, and argued that ductivity of the composite was significantly increased and the phase
the enhanced heat conduction compensated the suppressed natural change process was greatly enhanced.
convection, yielding significant enhancement in phase change rate. The above literature survey demonstrates that the influence of open-
Zhang et al. [23] conducted numerical simulations on similar scenario cell metal foam as heat spreader on phase change heat transfer has been
using a more realistic natural convection thermal boundary, with local mainly simulated using the VAM approach, and only heat conduction-
temperature difference between ligament surface and interstitial PCM dominated phase change was considered at pore-scale by the DNS ap-
measured at one selected point. Qu et al. [24] simulated the phase proach. The contribution of local natural convection at pore level to
change performance of a metal foam heat sink for power battery local phase interface evolution as well as global phase change behavior
thermal management. A two-dimensional (2D) model with local is yet justified, nor fully understood. Besides, systematical comparison
thermal non-equilibrium was adopted. Finally, it should be mentioned between the two approaches has not been performed. In the present
that the newly developed Lattice Boltzmann Method (LBM) was also study, melting heat transfer in open-cell metal foams was simulated
employed to investigate phase change heat transfer in porous materials using both the VAM and DNS approaches. Phase interface, temperature
[25–28]. and velocity fields obtained using the two approaches were system-
Although VAM is capable of predicting phase change behavior of atically compared. Particular attention was placed upon exploring the
porous media, it only provides an equivalent description of both fluid physical mechanisms underlying pore-scale phase change and justifying
flow and heat transfer characteristics in the complex foam structures, as the contribution of local natural convection to both microscopic (pore-
pore-scale features are lost under the basic assumption of this method. scale) and macroscopic (global) phase change behaviors. While the
For comparison, DNS has competitive preponderance in providing DNS-based results were used as benchmark of comparison, existing
physical insight into pore-scale phase change characteristics, such as experimental results were also used for better validation.
local solid-liquid interface shape and propagation, temperature

Fig. 1. Image of open-cell metal (copper) foam and reconstructed sphere-cut tetrakaidecahedron cells.

702
X. Yang et al. Applied Energy 229 (2018) 700–714

2. Direct numerical simulation (DNS) Table 1


Thermo-physical properties of PCM (Paraffin) [40] and aluminum [39].
2.1. Reconstruction of metal foam with open cells Material Parameter Value

High porosity open-cell metal foams typically have stochastic cel- Paraffin Melting temperature (K) 307 ∼ 309
Latent heat (J kg−1) 250,000
lular morphology, as shown in Fig. 1. To directly model the melting
Density (kg m−3) (solid/liquid) 880/760
heat transfer of PCM saturated in a metal foam block, the full size of the Special heat capacity (J kg−1 K−1) (solid/liquid) 2000
block should be considered. To simplify the problem, two fundamental Thermal conductivity (W m−1 K−1) 0.4/0.2
assumptions need to be made: one is to assume the stochastic metallic Kinematic viscosity (m2 s−1) 2.508 × 10−3
matrix to be periodic and the other is to employ tetrakaidecahedron Thermal expansion coefficient 0.001
Aluminum Density (kg m−3) 2719
unit cell (UC) to represent the real foam topology. Fig. 1 presents the
Special heat capacity (J kg−1 K−1) 871
reconstructed tetrakaidecahedron cells, which was centrally cut by a Thermal conductivity (W m−1 K−1) 202.4
sphere of 5.08 mm in diameter, resulting in a pore density of 5 PPI
(pores per inch). Correspondingly, the volume fraction (porosity) of the
UC (or foam) was 0.90. The UC of Fig. 1 was validated by our previous 2.4. Governing equations
DNS results on solidification of PCM saturated in metal foam [38].
For direct numerical simulation, heat transfer in metallic matrix was
2.2. Computational domain considered from heat transfer during phase change. The metallic matrix
was motionless, thus only energy equation was needed:
The periodicity of the problem posed above suggested that only one ∂Ts
piece of the whole structure was needed for direct numerical simula- ρs cps = ks ∇2 Ts
∂t (1)
tion, as illustrated in Fig. 2, with a total of 18 cells (3 × 6 × 1 for
x × y × z direction) considered to save computational cost. While all Conjugate heat transfer was considered at the interface between
tetrahedral elements were applied according to the geometric feature of PCM and ligament surface, as:
tetrakaidehedron cell, fine tetrahedral cells were applied on matrix
∂Tf ∂T
surface (Fig. 2) to catch the details of solid ligament surfaces. Tf = Ts, kf → = ks →s
∂n ∂n (2)

2.3. Initial and boundary conditions where ρ, cp and k were the density, specific heat and thermal con-
ductivity; the subscripts f and s stood for phase change material and
The front and rear faces of the computational domain were treated solid ligament, respectively.
as symmetry, and the top, the bottom as well as the right faces were Upon hating, the PCM in solid phase was going to melt, followed by
thermally insulated. The temperatures of the PCM and the metallic natural convection in the melt. The thermo-physical properties of the
matrix were initially assumed to be the same as Ti (< Tm1) that equaled PCM were assumed to be independent of temperature, except for the
to ambient air temperature. A constant temperature wall boundary was density of PCM in liquid phase (Table 1) [39,40]. The Boussinesq model
applied on the left face, enabling phase change in the immediate vici- was applied to the momentum equations to estimate the contribution of
nity of the heating wall. The initial temperature was set at 300 K and local free convection to the melting process. The governing equations
the heating wall temperature at 368 K. were given by:

Fig. 2. Computational domain and representative mesh for DNS (direct numerical simulation).

703
X. Yang et al. Applied Energy 229 (2018) 700–714

Continuity equation:

∇· U = 0 (3)

Momentum equation:

∂U → → → (1−f )2 →
ρf + ρf (U ·∇) U = −∇P + μf ∇2 U + ρf →
g β (Tf −Tm) + 3 l Am U
∂t fl + δ
(4)
Energy equation:

∂Tf → ∂f
ρf cpf + ρf cpf U ·∇Tf = ∇ ·(kf ∇Tf )−ρf L l
∂t ∂t (5)

where ρf →
g β (Tf −Tm) was introduced to account for the buoyancy flow in
(1 − f )2 →
the melt phase, 3 l Am U was introduced to damp the velocity in
fl + δ
Momentum equation within the solid PCM domain ( fl = 0 ), two nu-
merical constants δ and Am were the damping coefficient: one relatively
small (δ = 10−10 ) and the other sufficiently large ( Am = 1015 ); the last
term in Energy equation described the latent heat absorption during
phase change process and fl was the melting fraction. The melting
fraction was determined depending upon the melting temperature
range (Tm1 ∼ Tm2), as:

⎧ 0 at T ⩽ Tm1 solid
⎪ Tf − Tm1
fl = at Tm1 < T < Tm2 mushy
⎨ Tm2 − Tm1
⎪ 1 at T ⩾ Tm2 liquid (6) Fig. 3. Computational domain and representative mesh for simulation based on

volume averaging.

2.5. Numerical procedure

The commercially available software ANSYS-FLUENT 17.1 was


employed to solve the governing equations, with the finite volume
method (FVM) with tetrahedral elements adopted for the modeling.
High resolution scheme was selected for better computation accuracy
and the solution was thought to be converged when the residuals of all
the governing equations were less than 10−6. Three sets of meshes with
2,675,390, 5,341,442 and 8,851,256 elements and three time steps of
0.1 s, 0.05 s and 0.01 s were tested for the case of Ti = 300 K and
Tw = 368 K. It was established that the predicted values from the last
two sets of meshes exhibited a deviation less than 1.3%. Hence, the set
of mesh having the order of magnitude of 5,341,442 elements and a
time step of 0.05 s was used in subsequent simulations. Validation of Fig. 4. Dependence of temperature at fixed position on time: comparison of
the numerical simulation was discussed in Section 4. VAM-based simulation results with experimental measurements [23].

Table 2
3. Volume averaged method (VAM)
Comparison of numerical prediction and experimental measurement [23] at a
given points.
3.1. Treatment of PCM-foam composite layer
Time Texpt at TNS at Relative Texpt at TNS at Relative
elapsed/s ligament/K ligament/K error/% PCM/ PCM/ error/%
Porous media saturated with fluid have been commonly treated as
K K
equivalent fluid with effective thermophysical properties. The bulk
porous medium is divided into small control volumes consisted of one 1000 307.3 305.8 0.49 306 305.3 0.23
entire pore structure. Thermal transport at pore scale is then lumped 2000 315.8 313.7 0.67 314.1 313.1 0.32
3000 322.2 319.7 0.78 320.5 319.1 0.44
into each control volume where volume-averaged integral calculation is
4000 327.8 325.3 0.76 326.1 324.9 0.37
performed. With VAM, the heat and mass transfer in porous media can 5000 332.8 329.4 1.02 331.1 329.2 0.57
be solved by adding source terms to the original Navier-stokes and 6000 337.8 335.6 0.65 336.2 335.4 0.24
energy equations. 7000 342.2 340 0.64 341.2 339.9 0.38
The governing equations for melting phase change heat transfer in a 8000 345.5 342.8 0.78 344.7 342.8 0.55
9000 348.2 345.4 0.80 347.8 345.3 0.72
PCM-metal-foam composite layer can be summarized as follows:
10,000 351 349.6 0.40 351.2 349.6 0.46
Continuity equation:
Note: Expt and NS stand for experimental measurement and numerical simu-
∂ρf →
+ ∇ (ρf 〈U 〉) = 0 lation, in respective.
∂t (7)

Momentum equations:

704
X. Yang et al. Applied Energy 229 (2018) 700–714

Fig. 5. Comparison between solid-liquid interface evolutions obtained separately with VAM and DNS: (a) with local natural convection considered; (b) conduction
dominated (natural convection neglected).

705
X. Yang et al. Applied Energy 229 (2018) 700–714

density difference between solid and liquid phases was neglected. The
open-cell metal foam of Fig. 1 was taken as homogeneous and isotropic
with respect to its effective thermophysical properties.

3.2. Thermophysical properties

To solve the governing equations for melting heat transfer in PCM-


saturated open-cell metal foams, the structural parameter (asf ), the in-
terstitial heat transfer coefficient (hsf ) as well as the key thermophysical
properties (ke, kfe, kse, ktd, K , CE ) needed to be determined. As pre-
viously mentioned, for the VAM, pore-scale structural features were
lumped into the macroscopic parameters, e.g., the porosity and pore
density. Calmidi et al. [41] proposed a model to link pore-scale struc-
Fig. 6. Melting fraction plotted as a function of time: comparison between VAM tural parameters to the macroscopic ones, and obtained the specific
and DNS. area as:
1.18
ρf ∂〈u〉 ρf → ∂〈p〉 μf μf ρf CE → ⎞ asf = 3π (1−ε )
+ 2 (〈U 〉·∇) 〈u〉 = − + ∇2 〈u〉−⎜⎛ + |〈U 〉|⎟ 〈u〉 dp (13)
σ ∂t σ ∂x σ ⎝ K K ⎠
(1−fl )2 where dp denoted the average pore size of open-cell foam and could be
− Am 〈u〉 estimated by:
fl3 +δ (8)
0.0224
ρf ∂〈v〉 ρf → μf μf ρf CE → ⎞ dp =
∂〈p〉 ω (14)
+ 2 (〈U 〉·∇) 〈v〉 = − + ∇2 〈v〉−⎜⎛ + |〈U 〉|⎟ 〈v〉
σ ∂t σ ∂y σ ⎝K K ⎠ The pore density ω was commonly provided by the manufacturer.
(1−fl )2 The interstitial heat transfer coefficient could in general not be di-
− Am 〈v〉 + ρf gβ (〈Tf 〉−Tm1)
fl3 + δ (9) rectly measured by experiments due to the complex porous structure
and the complicated fluid flow condition within a pore. Consequently,
ρf ∂〈w〉 ρf → ∂〈p〉 μf μf ρf CE → ⎞ the empirical model of Žukauskas [42] was often employed to simplify
+ 2 (〈U 〉·∇) 〈w〉 = − + ∇2 〈w〉−⎜⎛ + |〈U 〉|⎟ 〈w〉
σ ∂t σ ∂z σ ⎝ K K ⎠ the heat transfer scenario in open-cell foams and satisfactory agreement
(1−fl )2 was achieved. In the present study, this model was also employed, as
− Am 〈w〉 [42]:
fl3 + δ (10)
0.4 0.37
→ ⎧ 0.76Re Pr kf /d, 0 < Re⩽40
where the Forchheimer extended Darcy equation ( μf
K
+
ρf CE
K )
|〈U 〉| 〈u〉 ⎪
hsf = 0.52Re0.5Pr 0.37kf / d, 40 < Re⩽1000
was employed to describe the additional flow resistance induced by ⎨
(1 − fl )2 ⎪ 0.26Re0.6Pr 0.37kf / d, 1000 < Re⩽20000 (15)
metal foam, Am 〈u〉 denoted the source term for damping the ve- ⎩
fl3 + δ
locity before melting occurred (with numerical constants taking same where d was the average ligament diameter of the open-cell foam. In
values as those for DNS). The free motion driven by temperature dif- practice, d could be estimated by examining scanning electron micro-
ference in the melt phase was assumed to be Boussinesq free convec- scope (SEM) images.
tion, governed by ρf gβ (〈Tf 〉−Tm1) in Momentum equation. With regard to thermophysical properties, the following models
For the energy equations, due to large difference in thermophysical were employed.
properties between metal ligaments and PCM, thermal non-equilibrium Effective thermal conductivity [43]:
was considered to model transient melting heat transfer between liga-
ke (1−ε ) kf
ments and PCM. Therefore, the two-temperature model was adopted, = + ε
as: ks
(
1−e +
3e
2α ) ⎡⎣3(1−e) + 3
2
αe⎤

ks
(16)
Energy equation for PCM:
Thermal dispersion conductivity [44]:
df ∂〈Tf 〉 →
ερf ⎛cpf + L l ⎞
⎜ ⎟ + ρf cpf 〈U 〉·∇〈Tf 〉 = ∇ ·((kfe + ktd ) ∇〈Tf 〉) 0.36
⎝ dt ⎠ ∂t ktd = ρ cpf d u2 + v 2 + w 2
1−ε f (17)
−hsf asf (〈Tf 〉−〈Ts 〉) (11)
Permeability and inertial coefficient [45]:
Energy equation for metal foam:
ε [1−(1−ε )1/3]
∂〈Ts 〉 K= dp2
(1−ε ) ρs cps = ∇ ·(kse ∇〈Tf 〉)−hsf asf (〈Ts 〉−〈Tf 〉) 108[(1−ε )1/3−(1−ε )] (18)
∂t (12)
where ∇ ·((kfe + ktd ) ∇〈Tf 〉) in Energy equation for PCM (Eq. (11)) de- −1
cd ε ⎛1.18 1−ε 1 ⎞
scribed the thermal diffusivity term and here the thermal dispersion CE = 0.095 ⎜ ⎟
12 3(χ −1) ⎝ 3π G ⎠ (19)
conductivity ktd was also accounted for in the melt phase; while in
Energy equation for metal foam (Eq. (12)) only thermal conductivity of In the above equations, ke, kf, ks denoted the (effective) thermal
metal foam was considered; for both two Energy equations (Eqs. (11) conductivity of PCM-metal foam composite, saturating fluid and solid
and (12)), hsf asf (〈Ts 〉−〈Tf 〉) denoted the interstitial heat transfer be- ligaments; e and α were the thickness and cross-sectional area ratio of
tween solid ligaments and the saturating PCM. node to solid ligament (e = 0.3, α = 1.5); cd was the drag coefficient
To numerically simulate the transient phase change behavior of which was suggested to have a mean value of 1.56 by Bhattacharya
PCM saturated in open-cell metal foam, a few assumptions were also et al. [46] in the porosity range of 0.8991–0.978; χ and G represented
made for simplification. PCM in liquid phase was considered to be in- the flow tortuosity and shape function for metallic ligaments, which
compressible Newtonian fluid and the volume expansion of PCM due to could be calculated by:

706
X. Yang et al. Applied Energy 229 (2018) 700–714

ε 3.3. Computational domain and initial/boundary conditions


χ=
1−(1−ε )1/3 (20)
For VAM-based numerical simulations, the PCM-metal foam com-
G = 1−e−(1 − ε )/0.04 (21) posite was treated as an equivalent fluid having effective thermo-
physical properties. The metallic ligaments and interstitial fluid were

Fig. 7. Temperature distribution of PCM on the mid-plane (z/t = 1/2) at Fo = 0.0057, 0.0113, 0.0171, 0.0227: comparison between VAM and DNS.

707
X. Yang et al. Applied Energy 229 (2018) 700–714

not distinguished physically in the computational domain, which had Technique were employed for numerical computation. During each
the size of L × t × H (length × width × height) as shown in Fig. 3. The iteration, the convergence was obtained only if the residual was less
initial conditions were: than 10−6.
The solution accuracy for a transient problem was significantly in-
〈Tf 〉 = 〈Ts 〉 = Ti, at t = 0 (22)
fluenced by mesh quality and time step. Both the mesh and time step
The boundary conditions were: sensitivities were tested using three sets of meshes with 80 × 100 × 15,
110 × 140 × 30 and of 130 × 150 × 45 elements and three sets of
〈Tf 〉|x = 0 = 〈Ts 〉|x = 0 = Tw (23)
time steps with 0.1 s, 0.05 s and 0.01 s. The initial and boundary con-
∂〈Tf 〉 ∂〈Ts 〉 ditions were previously presented: Ti = 300 K, Tw = 368 K, the right,
= =0 top and bottom faces were thermally insulated and the other two walls
∂x ∂x x=L (24)
x=L
were symmetry planes. The test sequence was as follows: first, with a
∂〈Tf 〉 ∂〈Ts 〉 ∂〈Tf 〉 ∂〈Ts 〉 time step of 0.5 s, three sets of meshes of 80 × 100 × 15,
= = = =0 110 × 140 × 30 and of 130 × 150 × 45 were tested; then, with a fixed
∂y y=0
∂y y=0
∂y y=H
∂y y=H (25) mesh density of 110 × 140 × 30, three time steps of 0.5 s, 0.1 s and
∂〈Tf 〉 ∂〈Tf 〉 0.05 s were tested. By comparing the numerical results obtained with
∂〈Ts 〉 ∂〈Ts 〉
= = = =0 different meshes and time steps, it was established that the mesh with a
∂z z=0
∂z z=t ∂z z=0
∂z z=t (26) total element size of 110 × 140 × 30 and a time step of 0.05 s was
sufficient to achieve high accuracy, as the predicted values from the last
3.4. Sensitivity test two sets of meshes/time steps exhibited a deviation less than 1.7%.

To solve the transient heat conduction with latent heat released and 4. Validation
the local natural convection in melt phase, the finite volume method
(FVM) was employed. The Pressure-Linked Equations algorithms Both the DNS and VAM approaches were used to characterize
(known as SIMPLE) was employed for pressure and velocity coupling. transient melting heat transfer in PCM saturated in open-cell metal
The staggered grid system with structural meshes was applied to dis- foam.
cretize the continuity, momentum and energy equations. The Tri For the DNS approach, melting phase change occurred in the PCM
Diagonal Matrix Algorithm (TDMA) and the Gauss-Seidel Iteration trapped in individual pores and heat transfer occurred between metallic

Fig. 8. Flow streamlines at selected time at Fo = 0.0057, 0.0113, 0.0171 and 0.0227: comparison between VAM and DNS.

708
X. Yang et al. Applied Energy 229 (2018) 700–714

Fig. 9. DNS-predicted temperature distributions of PCM on selected planes (z/t = 0, 1/6, 1/3 and 1/2) at Fo = 0.0171.

ligaments surfaces and the trapped PCM. If the pore-scale structural computational models developed in the present study. Further, by
features of the foam were precisely captured, the phase change beha- employing the non-equilibrium thermal model, slight difference be-
vior of the PCM-foam composite could be accurately characterized. In tween the ligament surface and the saturating paraffin was also ob-
the present study, inspired by our previous studies on heat conduction served in Fig. 4.
in open-cell foams [38], Tetrakaidehedron cells were again used to
reconstruct the topological features of the foam.
As for the VAM-based numerical model, significant assumption was 5. Results and discussion
made: the metal-foam-PCM composite was approximated as an
equivalent fluid that had the same effective thermophysical properties 5.1. Solid-liquid interface
as the composite. The sequence for model validation was as follows:
first, the VAM-based simulation was validated by comparing the pre- Fig. 5 compared the evolution of solid-liquid interface predicted
dicted temperature variation history with existing experimental mea- using the VAM with that using the DNS at four selected times
surements [23] under the same initial and boundary conditions ad- Fo = 0.0057, 0.0113, 0.0171 and 0.0227. Particular effort was made to
dressed in Ref. [23]; subsequently, the DNS-based simulation was justify the contribution of local natural convection to the evolution. Due
further validated by quantitatively comparing with the VAM-based si- to the symmetry nature in VAM simulations, only the 2D results were
mulation results. Zhang et al. [23] conducted experimental measure- depicted. For the DNS-based results, the grey and white colors denoted
ment on the transient melting heat transfer in paraffin embedded in separately the solid and liquid paraffin, and the tetrakaidehedron cells
open-cell copper foam under the constant heat flux thermal boundary. in brown color could be clearly seen. The blue and red colors in the
The rectangular enclosure with the inner size of VAM-based results set the two melting fraction limits, 0 and 1.0, which
100.0 × 100.0 × 10.0 mm was used to house the metal foam saturated represented the paraffin in solid phase and liquid phase respectively;
by paraffin (composite PCM). The center of the composite PCM was and the other colors turning from blue to red denoted the mushy zone
selected as the point for monitoring temperature variation. Two ther- (mixture of solid and liquid phases). Due to the narrow temperature
mocouples were employed to measure the temperatures for metallic range of phase change (307 ∼ 309 K) in the present PCM, the thickness
ligament surface and the PCM inside the pore space, in respective. One of the mushy zone was not large. Further, given the fundamental as-
was directly inserted in the pore space filled with PCM and the other sumption underlying the VAM approach, the metallic ligaments could
was soldered on the ligament surface. Fig. 4 compared the temperature not be seen in the simulation results.
variation history at selected points obtained from the VAM approach Fig. 5(a) demonstrated the specific features of solid-liquid interface
with the experimental measurements [23]. Good agreement was evolution. Initially, once the solid paraffin adjacent the heating
achieved between numerical and experimental data (maximum devia- boundary started to melt, the interface gradually moved away from the
tion 1.02%, see Table 2), confirming the feasibility of the boundary. At this stage, the volume of the melt phase was so small that
natural convection was not strong enough to enhance local heat

709
X. Yang et al. Applied Energy 229 (2018) 700–714

transfer. Hence, the solid-liquid interface was globally flat (see


Fig. 5(a), DNS result at Fo = 0.0057) and parallel to the heating
boundary. Subsequently, more solid paraffin was melt and the melt
phase flew upward, leading to local acceleration of interface evolution
(see Fig. 5(a), DNS results at Fo = 0.0171 and 0.0227). This contributed
to a macroscopically “slope” shape of the phase interface. From pore-
scale point of view, it could be seen that the melting interface was
microscopically uneven in contrast to the smooth one observed in VAM
simulations (Fig. 5(a)). A large portion of the PCM adjacent the metallic
ligaments was able to be melt at a faster speed, and the PCM located in
the bigger window of the tetrakaidehedron cell was still in solid phase
while the solid PCM next to the ligaments had already melt (Fig. 5(a),
DNS results). In contrast, with VAM simulations, no metallic ligaments
could be observed and a continuous melting interface separated the
solid and liquid paraffin. However, broadly speaking, the two numer-
ical models yielded qualitatively similar predictions of transient solid-
liquid interface evolution.
The DNS results of Fig. 5(a) and (b) also quantified the contribution
of local natural convection to melting interface propagation. In com-
parison, the solid-liquid phase interface predicted by the VAM was
globally flat and parallel to the heating boundary due to the conduc-
tion-dominated heat transfer mechanism. Initially, the volume of melt
phase was tiny so that the phase interface was similar, with or without
natural convection considered (see Fig. 5(a) and (b), the DNS results).
With time elapsed, local natural convection was gradually enhanced
which accelerated interface evolution in the top domain (see Fig. 5,
DNS results at Fo = 0.0227). However, the conduction-dominated heat
transfer mode resulted in a globally flat solid-liquid phase interface.
This phenomenon was also observed in the VAM-based results without
natural convection considered.
The contribution of natural convection to the evolution of melting
interface was addressed qualitatively in Fig. 5. To quantitatively com-
pare the two models, the melting fraction curve was plotted in Fig. 6.
The DNS approach predicted an increasing trend of the melting fraction
as a function of dimensionless time (Fo number), quantitatively similar
to that predicted by the VAM approach. This provided an indirect va-
lidation of the present pore-scale numerical model (i.e., the DNS ap-
proach). According to the development of local natural convection, the
history of interface evolution could be divided into three stages: the
pre-developed stage, the developed stage, and the post-developed stage.
At the pre-developed stage, the curve with natural convection con-
sidered overlapped with the conduction-dominated one; when more
solid paraffin was melt (melting fraction exceeding 0.3), the difference
between the two cases gradually increased, which was consistent with
the qualitative observation in Fig. 5. This could be explained as follows:
solid paraffin adjacent the heating boundary initially started to melt but
the melting phase had such a small volume that strong natural con-
vection was not developed, thus limiting the heat transfer enhancement
by natural convection. At the developed stage, more solid paraffin was
melt and the gradually-increased melting volume enabled convection
development. This could be validated by the interface evolution results
of Fig. 5 where the interface morphology gradually changed from plane
to slope. At this stage, natural convection played a vital role in en-
hancing the phase change heat transfer. Beyond this stage, the solid
paraffin formed a curve-edged triangle located in the right corner of the
studied domain, resulting in the post-developed stage. The solid PCM in
this corner region took about 30% of the full melting time to melt.
The present results demonstrated that while the VAM approach was
Fig. 10. Comparison between VAM- and DNS-predicted temperature profiles at
select points: (a) natural convection dominated case; (b) heat conduction
capable of predicting the overall phase change behavior both qualita-
dominated case; temperature variations at selected points on metallic ligaments tively and quantitatively, the pore-scale melting features were lumped
and interstitial PCM with natural convection (c) and without natural convection into the homogenization treatment of the PCM-foam composite. In
(d). comparison, the DNS approach provided not only good prediction of
the overall phase change behavior but also captured pore-scale phase
change features. Nevertheless, the computational time needed by the
DNS approach was 29.4 times longer than that needed by the VAM
approach for the same phase change process.

710
X. Yang et al. Applied Energy 229 (2018) 700–714

presence of metallic ligaments resulted in locally tortuous isothermal


lines; for volume-averaged simulation, the isothermal lines resembled
the melting interface shown in Fig. 5. With natural convection con-
sidered, the uniformity of temperature in the gravity direction (y axis)
vanished. In contrast, when conduction was dominant, the temperature
was more uniform (isothermal lines more parallel to y-axis). On the left
surface, the PCM adjacent the heating boundary as well as the top re-
gion exhibited a higher temperature than PCM in other regions, due
mainly to the significant contribution of local natural convection to
melting heat transfer. This could be further validated by streamlines in
the melt phase shown in Fig. 8. A maximum flow velocity of 0.01 m/s
was found near the heating boundary and natural convection became
increasingly notable as more PCM was melt. Local natural convective
circulation was qualitatively predicted by the two numerical ap-
proaches. The fluid (melt) washed the solid PCM from the top to the
bottom, leading to the “slop” shape of the solid-liquid phase interface.
At pore scale, due to the presence of metal ligaments, the flow lines
were considerably tortuous, which might explain the uneven melting
interface observed in the vicinity of these ligaments.
For better demonstration of the pore-scale features of local tem-
perature distribution, comparisons were made among different planes
(z/t = 0, 1/6, 1/3 and 1/2) at Fo = 0.0171. As could be seen from
Fig. 9, natural convection played a significant role in shifting the
temperature distribution within the top region of the computation do-
main. The temperature fields globally resembled each other at different
planes but differed locally near the ligaments. The ligaments helped
increase the PCM temperature nearby. To quantitatively illustrate the
temperature variation history, two points located in the PCM domain
Fig. 11. (a) Comparison between DNS- and VAM-predicted boundary wall heat were selected to demonstrate the cases with/without natural convec-
flux profiles; (b) DNS-predicted average Nu number, both with and without tion, as shown in Fig. 10(a) and (b). Both approaches gave quantitative
natural convection considered. predictions of temperature variation. The temperature at P1 behaved
similarly in trend as that at P2 but differed in values for the case with
5.2. Temperature distribution natural convection. Comparatively, both the trend and value for tem-
perature at P1 and P2 were the same when natural convection was
Fig. 7 presented snapshots of temperature distribution on the mid- neglected. This might serve as a quantitative evidence for the uniform
plane (z/t = 1/2) at selected times (Fo = 0.0057, 0.0113, 0.0171, temperature distribution along gravity direction (y axis) in Figs. 7 and
0.0227). Again, two different approaches (DNS and VAM) contribution 9.
were employed. The temperature distribution was consistent with the To show local thermal non-equilibrium, temperature variation his-
melting interface evolution results of Fig. 5. For instance, abundant tories of the ligament surface as well as the saturating PCM were illu-
“slope” isothermal lines were observed: for pore-scale simulation, the strated in Fig. 10(c) and (d). Another five points, P3 to P7, were

Fig. 12. Schematic illustration of TES unit assemblies.

711
X. Yang et al. Applied Energy 229 (2018) 700–714

Table 3 Nu number with natural convection considered was 7.7 times higher
Item used for profit estimation. than that without. This indicated, from the heat transfer point of view,
Cut/$·m−3 V/m−3 η /% N/h tfull/s ε Ch/$·GJ−1 Π/$·yr−1 that natural convection should be given sufficient justification when
dealing with melting phase change process. However, although the Nu
750 0.03 5 10 146 (MF) 0.97 6.67 3440 (MF) ratio (7.7) of the convection-dominated case to the conduction-domi-
200 2580 nated one was large, the full melting time was only reduced by 45.5%.
(PF) (PF)
This might be explained as follows: initially there was insufficient vo-
Note: MF and PF stand for metal foam and plate fin, in respective. lume of the melt phase so that natural convection occurred locally near
the heating boundary, which was gradually enhanced after the melting
selected on the x−z plane at y/H = 0.5: P3 was located in the PCM fraction reached 0.40 (the two curves of Fig. 6 were coincident when
domain in the middle of the square cross-section of the tetra- the melting fraction was lower than 0.40). Further, in view of the fluid
kaidehedron cell; the other four points, P4 to P7, were located in the motion features associated with natural convection, the top region al-
metallic ligament domain. Again, natural convection was considered in ways had fresh fluid with a higher temperature, thus enjoying a higher
Fig. 10(c) while heat conduction dominated the process in Fig. 10(d). local heat transfer rate (see temperature distribution in Fig. 7, VAM
As could be seen, there did exist temperature difference between liga- results with natural convection). On the other hand, the corner region
ment surface and interstitial PCM. The temperature in the PCM domain that was at the bottom and far from the heating boundary suffered
was initially lower than that on the ligament surface. However, after significantly from low heat transfer rate. This was in consistence with
melting occurred on y/H = 0.5, i.e., Fo > 0.227, the PCM temperature the results of melting interface propagation in Figs. 5 and 9. As a result,
became equal to that on ligament surface. The initially lower tem- for the solid PCM in the corner region to melt, about 30% of the full
perature in the PCM domain was due mainly to the fact that the PCM melting time was consumed. During this period of time (i.e., the post-
absorbed heat and changed phase but maintained a smaller tempera- developed stage), the contribution of natural convection to melting heat
ture rise during phase change. Upon phase changing, the temperature transfer was not as much as that in the “developed” stage. However, by
difference gradually decreased, eventually approaching local thermal observing the melting interface evolution history (Fig. 5(b), the VAM
equilibrium at pore scale. results) in the conduction-dominated case, the interface was uniform in
the gravity direction and the corner area was smaller than that with
5.3. Heat transfer performance natural convection. Therefore, the full melting time was only reduced
by 45.5%, rather than 7.7 times less.
Numerical simulation can provide plenty of results in temperature
and velocity distribution within the whole computational domain 6. Techno-economic assessment
during the full melting process. To characterize the heat transfer per-
formance of phase change process, some advanced post-processing will 6.1. Energy recovery from waste heat
be carried out in the following part.
To quantify the effect of natural convection on melting heat Waste heat is inevitably produced during an industrial process and
transfer, a modified Nusselt number based on the heat transfer coeffi- released by two main exhausts: gas or water. Industrial process pro-
cient of the heating wall was introduced, as: duces a large amount of waste heat. For example, the cooling demand of
both industrial sectors and buildings in Europe is about 2.21 × 109 GJ/
h (t ) D q (t ) D y, and 78% of the total cooling demand is generated by industrial
Nu (t ) = = w
kf (Tw−Ti ) kf (27) cooling applications (1.72 × 109 GJ/a) [47]. Direct release of waste
heat to environment not only reduces the overall energy efficiency of an
where h (t ) was the transient heat transfer coefficient, D was the char-
industrial process, but also makes thermal pollution to the environ-
acteristic length of metal foam (thickness L), kf was the thermal con-
ment. Outdoor thermal comfort level is significantly degraded due
ductivity of the liquid PCM, qw(t) was the surface heat flux on the
mainly to the increased outdoor air temperature [48]. To both increase
heating boundary (which was the area-averaged value for DNS), Ti and
the energy efficiency and avoid polluting the environment, WHR
Tw were the temperatures of the initial state and heating wall, respec-
technologies seem to be an appropriate way. WHR collects, stores and
tively. With both DNS and VAM, the predicted Nusselt number was
transforms waste heat to useful thermal energy, either directly to the
plotted as a function of time in Fig. 11(a). There was good agreement
thermodynamics process or provide hot water and central heating for
between the two approaches, with or without considering natural
residential buildings. By using WHR, on one hand, the energy efficiency
convection. Thus, from the heat transfer point of view, these results
of an industrial process is improved; on the other hand, the energy
clearly demonstrated the feasibility and flexibility of the present pore-
consumption of fossil fuels to provide residential heating can be sig-
scale simulations. Consideration of natural convection enhanced sig-
nificantly reduced, with less carbon dioxide, oxynitride and oxysulfide
nificantly the Nusselt number. For example, the Nu number with nat-
emissions.
ural convection was 11.4 times higher than that without natural con-
vection at Fo = 0.02.
6.2. Technical assessment of metal foam composite PCM
To compare transient quantities during phase change, Yang et al.
[38] recommended an integral averaged method over the entire phase
Metal foams have promising structural controllability and robust-
change process. With ϕ (t ) and tfull denoting the transient quantity and
ness for they can be cut into pieces and sintered with heat transfer
the full phase change time, the integration was performed as:
environment as channel or tube heat exchanger in typical engineering
t
∫0 full ϕ (t ) dt applications [38]. Metal-foam-cored heat exchanger can be directed
ϕ = used as the key element for a WHR system, favoring a high heat transfer
t full (28)
efficiency. Besides, the metal foam/PCM composites can be packed
The, upon integrating upon time, the average Nu number (Nu ) separately and get charging/discharging like battery to form a flexible
could be obtained as: energy supplier. It is very convenient to assemble several composite
tfull qw (t ) D PCM in parallel or series, to get either a much larger TES capacity or a
Nu = ∫0 (Tw−Ti ) kf t full
dt
(29)
cascaded utilization of energy, as shown in Fig. 12. The present study
not only puts a physical insight into the thermo-fluidic characteristic
The corresponding results were presented in Fig. 11(b). The average into the transport phenomena in porous media with interstitial phase

712
X. Yang et al. Applied Energy 229 (2018) 700–714

change occurred, but also provides heat transfer performance evalua- the economic applicability of metal foam as heat transfer enhancement
tion of composite PCMs potentially for WHR. Practical design of WHR for TES. It is pleasantly expected to obtain much more profits if large
heat exchanger may benefit from the analysis of phase change heat scale of the metal-foam-cored plate heat exchanger was used for TES.
transfer in composite PCM unit.
7. Conclusions
6.3. Preliminary economic analysis
The melting behavior of phase change material (PCM) impregnated
Open-cell metal foam can significantly enhance phase change heat into open-cell metal foam was investigated from both macroscopic and
transfer, favoring a higher heat transfer coefficient. A higher heat microscopic point of view using two different approaches: the volume-
transfer coefficient indicates that the temperature difference between averaged method and the unit-cell-based direct numerical simulation.
the heat transfer media in two sides (exhaust gas and PCM) is smaller; The effects of natural convection on solid-liquid interface propagation,
i.e. WHR system using metal foam for TES heat exchanger can extract local and global temperature distributions and wall heat transfer coef-
thermal energy from exhaust gas with a lower temperature. On the ficient were justified, both qualitatively and quantitatively. Main con-
other hand, if the temperature for the exhaust gas is fixed, using metal clusions were drawn as follows:
foam is capable of tremendously reducing the full melting time for
thermal energy storage. This reveals that more waste energy can be (1) Temperature difference between metal ligaments surface and the
recovered and stored by the metal-foam cored phase change heat ex- saturating PCM was not negligible, thus demonstrating the applic-
changer. Let’s compare the lifetime cost of TES unit with plate fin and ability of the two-temperature model for modelling phase change in
using metal foam as heat transfer enhancement media. The capital in- PCM-metal-foam composites.
vestment Cci using metal foam contains two parts: material cost Cm and (2) The volume-averaged method predicted both qualitatively and
the welding cost Cw. Therefore, the capital investment can be expressed quantitatively similar phase change behaviors as the direct nu-
as: merical simulation from the global point of view. The direct nu-
merical simulation could additionally capture pore-scale features of
Cci = Cm + Cw (30)
solid-liquid interface evolution and temperature distributions in
Regarding that metal foam will be inserted and welded into the cellular foam structure as well as interstitial PCM.
space between two plates, the material cost is thus depended on the (3) The average Nu number calculated with natural convection con-
void fraction of the TES unit V (volume for PCM). According to the sidered was 7.7 times higher than that without considering natural
manufacturing routine, the welding cost is typically estimated to be η convection. The melting front in the former propagated faster than
percent of the total material cost. Therefore, the capital investment the latter, leading to 45.5% reduction in full melting time. These
turns to be results indicated the significant role played by buoyancy flow in
Cci = Cut V (1 + η) melting phase change process.
(31)
(4) For the same operating hours (10 h per day), metal-foam-cored TES
−3
where Cut ($·m ) is the specific cost the materials. unit can store more energy compared with the case of plate fin heat
The capital cost increases with using metal foam, while it can ex- exchanger. A payback time of 6.47 year is found to use metal foam
tract and store more energy from exhaust gas for a given period of time. for TES and a considerable amount of money (25, 800 $) can be
For comparison, two cases for TES with/without metal foam were si- obtained by using metal foam with a total 7.4 m3 TES tank, after a
mulated: one used plate fin style; the other one inserted metal foam into 30-year operating.
the spaces between every two plates. The full melting time for the TES
unit is tfull, and the available charging hours is N per day; therefore, the Acknowledgements
fully charged volume Vc can be estimated to be
N This work was supported by the National Natural Science
Vc = Vε Foundation of China (51506160), China Post-doctoral Science
t full (32)
Foundation Funded Project (2015M580845, 2016T90916), Shaanxi
Here, ε represents the porosity for metal foam and it equals to be unit Province Post-Doctoral Science Foundation Project (2016BSHYDZZ54),
for the case without metal foam. Compared with the amount of latent China Northwest Architecture Design and Research Institute CO. Ltd
heat, sensible heat stored in a TES unit within a medium-low tem- (TC-2016A-092), the fundamental research funds for central uni-
perature range can be neglected. The energy Qc stored in a TES unit can versities (xjj2016042) and the Beijing Key Lab of Heating, Gas Supply,
be then estimated as Ventilating and Air Conditioning Engineering (NR2016K01). Xiaohu
Qc = ρPCM Vc L Yang is financially supported by China Scholarship Council (CSC) to
(33)
conduct research in Royal Institute of Technology (KTH) and
If the stored thermal energy is to provide hot water or central Mälardalen University (MDH), Sweden.
heating for residential buildings, the profits Π for the case with metal
foam can be estimated to be References
N
Π = Ch Qc− Cci [1] Sharma SS. Determinants of carbon dioxide emissions: Empirical evidence from 69
t full (34) countries. Appl Energy 2011;88:376–82.
[2] Berardi U. Building energy consumption in US, EU, and BRIC countries. Proced Eng
where the prices for residential heating is Ch ($·GJ−1). 2015;118:128–36.
Table 3 shows the profit estimation of the practical TES unit with/ [3] Chua KJ, Chou SK, Yang WM, Yan J. Achieving better energy-efficient air con-
without metal foam. The profit for the waste heat to provide residential ditioning – A review of technologies and strategies. Appl Energy 2013;104:87–104.
[4] Hu Y, Yan J. Characterization of flue gas in oxy-coal combustion processes for CO2
hot water is calculated in one-year operation. Although metal foam capture. Appl Energy 2012;90:113–21.
material increases the capital cost (about $5,570 for total 7.4 m3) [5] Tian Y, Zhao L, Meng H, Sun L, Yan J. Estimation of un-used land potential for
compared with the plate fin case, more profits can be obtained due to biofuels development in (the) People’s Republic of China. Appl Energy
2009;86:S77–85.
more energy stored within the same operating hours. A 6.5 year is [6] Ji C, Qin Z, Dubey S, Choo FH, Duan F. Three-dimensional transient numerical study
found to be the payback time for using metal foam. For a 15-year on latent heat thermal storage for waste heat recovery from a low temperature gas
lifetime, the metal-foam-cored plate heat exchanger (7.4 m3) can save flow. Appl Energy 2017;205:1–12.
[7] Yang X, Lu Z, Bai Q, Zhang Q, Jin L, Yan J. Thermal performance of a shell-and-tube
about $12,900 in comparison with plate heat exchanger. This indicates

713
X. Yang et al. Applied Energy 229 (2018) 700–714

latent heat thermal energy storage unit: Role of annular fins. Appl Energy convection in porous media under local thermal non-equilibrium conditions. Int J
2017;202:558–70. Heat Mass Transf 2014;70:979–89.
[8] Yang X, Lu T, Kim T. Temperature effects on the effective thermal conductivity of [28] Tao YB, You Y, He YL. Lattice Boltzmann simulation on phase change heat transfer
phase change materials with two distinctive phases. Int Comm Heat Mass Transfer in metal foams/paraffin composite phase change material. Appl Therm Eng
2011;38:1344–8. 2016;93:476–85.
[9] Liu YD, Zhou YG, Tong MW, Zhou XS. Experimental study of thermal conductivity [29] Hu X, Patnaik SS. Modeling phase change material in micro-foam under constant
and phase change performance of nanofluids PCMs. Microfluid Nanofluid temperature condition. Int J Heat Mass Transf 2014;68:677–82.
2009;7:579–84. [30] Sundarram SS, Li W. The effect of pore size and porosity on thermal management
[10] Fan L, Khodadadi J. Thermal conductivity enhancement of phase change materials performance of phase change material infiltrated microcellular metal foams. Appl
for thermal energy storage: a review. Renew Sustain Energy Rev 2011;15:24–46. Therm Eng 2014;64:147–54.
[11] Zhao CY, Lu W, Tian Y. Heat transfer enhancement for thermal energy storage using [31] Feng SS, Shi M, Li YF, Lu TJ. Pore-scale and volume-averaged numerical simula-
metal foams embedded within phase change materials (PCMs). Sol Energy tions of melting phase change heat transfer in finned metal foam. Int J Heat Mass
2010;84:1402–12. Transf 2015;90:838–47.
[12] Zhao CY, Wu Z. Heat transfer enhancement of high temperature thermal energy [32] Hu X, Wan H, Patnaik SS. Numerical modeling of heat transfer in open-cell micro-
storage using metal foams and expanded graphite. Solar Energy Mater Solar Cells foam with phase change material. Int J Heat Mass Transf 2015;88:617–26.
2011;95:636–43. [33] Yao Y, Wu H, Liu Z. Numerical study of heat conduction of high porosity open-cell
[13] Fleming E, Wen SY, Shi L, da Silva AK. Experimental and theoretical analysis of an metal foam/paraffin composite at pore scale. ASME 2016 heat transfer summer
aluminum foam enhanced phase change thermal storage unit. Int J Heat Mass conference collocated with the ASME 2016 fluids engineering division summer
Transf 2015;82:273–81. meeting and the ASME 2016 14th international conference on nanochannels, mi-
[14] Beckermann C, Viskanta R. Natural-convection solid liquid-phase change in porous- crochannels, and minichannels: American society of mechanical engineers; 2016. p.
media. Int J Heat Mass Transf 1988;31:35–46. V001T02A6-VT02A6.
[15] Feng SS, Zhang Y, Shi M, Wen T, Lu TJ. Unidirectional freezing of phase change [34] Zhang Z, He X. Three-dimensional numerical study on solid-liquid phase change
materials saturated in open-cell metal foams. Appl Therm Eng 2015;88:315–21. within open-celled aluminum foam with porosity gradient. Appl Therm Eng
[16] Yang X, Wang W, Yang C, Jin L, Lu TJ. Solidification of fluid saturated in open-cell 2017;113:298–308.
metallic foams with graded morphologies. Int J Heat Mass Transf 2016;98:60–9. [35] Sedeh MM, Khodadadi J. Thermal conductivity improvement of phase change
[17] Harris KT, Haji-Sheikh A, Nnanna AA. Phase-change phenomena in porous media–a materials/graphite foam composites. Carbon 2013;60:117–28.
non-local thermal equilibrium model. Int J Heat Mass Transf 2001;44:1619–25. [36] Sedeh MM, Khodadadi J. Interface behavior and void formation during infiltration
[18] Mesalhy O, Lafdi K, Elgafy A, Bowman K. Numerical study for enhancing the of liquids into porous structures. Int J Multiphase Flow 2013;57:49–65.
thermal conductivity of phase change material (PCM) storage using high thermal [37] Sedeh MM, Khodadadi J. Solidification of phase change materials infiltrated in
conductivity porous matrix. Energy Convers Manage 2005;46:847–67. porous media in presence of voids. J Heat Transfer 2014;136. 112603.
[19] Krishnan S, Murthy JY, Garimella SV. A two-temperature model for solid-liquid [38] Yang X, Feng S, Zhang Q, Chai Y, Jin L, Lu TJ. The role of porous metal foam on the
phase change in metal foams. J Heat Trans 2005;127:995–1004. unidirectional solidification of saturating fluid for cold storage. Appl Energy
[20] Chen ZQ, Gu MW, Peng DH. Heat transfer performance analysis of a solar flat-plate 2017;194:508–21.
collector with an integrated metal foam porous structure filled with paraffin. Appl [39] Yang S, Tao W. Heat transfer. Beijing: Higher Education Press; 2004.
Therm Eng 2010;30:1967–73. [40] Rubitherm Technologies GmbH; 2017.
[21] Tian Y, Zhao CY. A numerical investigation of heat transfer in phase change ma- [41] Calmidi VV. Transport phenomena in high porosity fibrous metal foams. Boulder,
terials (PCMs) embedded in porous metals. Energy 2011;36:5539–46. CO: University of Colorado; 1998.
[22] Li WQ, Qu ZG, He YL, Tao WQ. Experimental and numerical studies on melting [42] Žukauskas A. Heat transfer from tubes in crossflow. Adv Heat Transf
phase change heat transfer in open-cell metallic foams filled with paraffin. Appl 1972;8:93–160.
Therm Eng 2012;37:1–9. [43] Yang XH, Bai JX, Yan HB, Kuang JJ, Lu TJ, Kim T. An analytical unit cell model for
[23] Zhang P, Meng Z, Zhu H, Wang Y, Peng S. Melting heat transfer characteristics of a the effective thermal conductivity of high porosity open-cell metal foams. Trans
composite phase change material fabricated by paraffin and metal foam. Appl Porous Med 2014;102:403–26.
Energy 2017;185:1971–83. [44] Georgiadis JG, Catton I. Dispersion in cellular thermal-convection in porous layers.
[24] Qu Z, Li W, Tao W. Numerical model of the passive thermal management system for Int J Heat Mass Transf 1988;31:1081–91.
high-power lithium ion battery by using porous metal foam saturated with phase [45] Yang XH, Lu TJ, Kim T. An analytical model for permeability of isotropic porous
change material. Int J Hydro Energy 2014;39:3904–13. media. Phys Lett A 2014;378:2308–11.
[25] Huo Y, Rao Z. Investigation of phase change material based battery thermal man- [46] Bhattacharya A, Calmidi VV, Mahajan RL. Thermophysical properties of high por-
agement at cold temperature using lattice Boltzmann method. Energy Convers osity metal foams. Int J Heat Mass Transf 2002;45:1017–31.
Manage 2017;133:204–15. [47] Oró E, Miró L, Farid MM, Martin V, Cabeza LF. Energy management and CO 2
[26] Gao DY, Chen ZQ. Lattice Boltzmann simulation of natural convection dominated mitigation using phase change materials (PCM) for thermal energy storage (TES) in
melting in a rectangular cavity filled with porous media. Int J Therm Sci cold storage and transport. Int J Refrig 2014;42:26–35.
2011;50:493–501. [48] Huang Q, Meng X, Yang X, Jin L, Liu X, Hu W. The ecological city: considering
[27] Gao DY, Chen ZQ, Chen L. A thermal lattice Boltzmann model for natural outdoor thermal environment. Energy Proce 2016;104:177–82.

714

You might also like