Sei sulla pagina 1di 13

International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

International Journal of Thermal Sciences


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

Lattice Boltzmann simulation of melting of a phase change material confined


within a cylindrical annulus with a conductive inner wall using a body-fitted
non-uniform mesh
Gholamreza Imani
Department of Mechanical Engineering, Persian Gulf University, 75168, Bushehr, Iran

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

Keywords: In this paper, the lattice Boltzmann method on a body-fitted non-uniform mesh is employed to simulate the
Melting outward melting of a pure phase change material (PCM) confined in a cylindrical annulus with a conducting
Body-fitted non-uniform mesh inner wall, for Ra = 10 4 − 105 and Ste = 1. To perform the propagation step on the non-uniform mesh, the Taylor
Taylor series least square-based lattice series least square method is used. After verification of the developed code against the experimental and nu-
Boltzmann method
merical benchmark solutions, the effects of Reynolds number, radius ratio rr , wall to PCM conductivity ratio λ ,
Concentric cylindrical annulus
dimensionless inner wall thickness t ∗ , and thermal boundary condition on the melting process are studied. It is
Conducting inner wall
concluded that the smaller radius ratios enhance the melting process. It is seen that the results of cases with
λ = 10 and λ = 100 are in good agreement with each other and with the corresponding cases where the wall
thickness is ignored. However, this is not the case for λ = 1. Moreover, it is seen that for λ = 10 and λ = 100 , the
bigger t ∗ enhances the melting process whereas, for λ = 1, the reverse is true. Finally, it is figured out that the
effect of applying the adiabatic conditions on the outer cylinder, instead of the constant temperature, enhances
the liquid fraction in such a way that this improvement is seen to be more pronounced for the times after the
melted PCM first touches the outer cylinder.

1. Introduction dominated melting of a PCM in a cylindrical horizontal annulus. Ba-


gheri et al. [9] employed the finite volume method to numerically si-
Latent heat thermal energy storage systems (LHTESS) with melting/ mulate the melting of a PCM with variable thermophysical properties in
solidification of the phase change materials (PCMs) have many en- a double-tube heat exchanger. Darzi et al. [10] used the FLUENT soft-
gineering applications. The examples are in solar energy and heat re- ware to numerically simulate the melting of a phase change material
covery systems, as the intermittent sources of energy, where LHTESS confined between the two horizontal cylinders in concentric and ec-
can play a significant role in filling the gap between the supply and centric arrangements. Many researchers also numerically simulated the
demand [1,2]. There are other examples of engineering applications of melting of a PCM in the LHTESS with shell and tube configuration
the LHTESS, including but not limited to, cooling of electronic devices [11–15]. A comprehensive review of the works that considered the
[3] and heating of buildings [4]. numerical simulation of the cylindrical LHTES systems can be found in
Numerical simulation of the LHTESS has received much attention the review papers written by Al-Abidi et al. [5], Agyenim et al. [6], and
during the past few decades [5]. Among different geometries of PCM Dhaidan and Khodadadi [16,17].
container investigated by researchers in the LHTESS, the cylindrical Nonetheless, all of the above-mentioned researchers used the con-
geometry accounts for nearly 70%, which can be used in different ventional numerical methods as well as the commercial software like
configurations such as a cylinder, a double pipe heat exchanger (or a FLUENT to simulate the melting/solidification of the PCM in the
cylindrical annulus), and the shell and tube system [6]. LHTESS. However, during the last few decades, the LBM has become a
Some examples of the papers aimed to numerically simulate the powerful and popular numerical tool for simulation of complex en-
cylindrical annulus LHTESS are as follows. Ho and Lin [7] used the gineering problems [18]. This popularity of LBM is mostly attributed to
finite difference method to simulate the outward melting of a phase its explicit algorithm, linear nature of the advection term, easy paral-
change material encapsulated in a horizontal cylindrical annulus. Ng lelization, easy coding, and ability to deal with the complex geometries
et al. [8] used the finite element method to investigate the convection- and phenomena [18].

E-mail address: g.r.imani@pgu.ac.ir.

https://doi.org/10.1016/j.ijthermalsci.2018.10.009
Received 14 December 2017; Received in revised form 5 September 2018; Accepted 7 October 2018
1290-0729/ © 2018 Elsevier Masson SAS. All rights reserved.

Please cite this article as: Imani, G., International Journal of Thermal Sciences, https://doi.org/10.1016/j.ijthermalsci.2018.10.009
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

Nomenclature t inner cylinder wall thickness (m)


t∗ dimensionless inner cylinder wall thickness,
fi pre-collision density distribution function t ∗ = t /[2(rio − rii )] (see Fig. 1)
f i∗ post-collision density distribution function V = (u, v ) velocity vector (m/s)
fieq equilibrium distribution function of fi {V }i unknown vector
{f }∗i known vector x,y Cartesian coordinates (m)
Fl local melt fraction x ∗, y∗ non-dimensional coordinates
gi pre-collision fluid internal energy distribution function
gi∗ post-collision fluid internal energy distribution function Greek symbols
gieq equilibrium distribution function of gi
gwi pre-collision wall medium internal energy distribution α thermal diffusivity (m2/2)
function δx lattice spacing
kf fluid thermal conductivity (W/mK) δt time step
kw wall thermal conductivity (W/mK) Θ dimensionless temperature, Θ = (T − Tm)/(Th − Tm)
Lf latent heat of the melting of the PCM (kJ / kg ) Λ wall to PCM thermal conductivity ratio, λ = k w / kf
Pr Prandtl number, ν / αf νf kinematic viscosity of the fluid (m2/2)
Ra Rayleigh number, Ra = gβ (Th − Tc )[2(ro − rio)]3 / ναf ρf density of the fluid (kg/m3)
rr radius ratio (ρcp) f volumetric heat capacity of the fluid (J/Km3)
[S ]i coefficient matrix (ρcp) w volumetric heat capacity of the wall (J/Km3)
[S ]Ti transpose of the coefficient matrix Σ wall to PCM volumetric heat capacity ratio,
Ste Stefan number, Ste = cp (Th − Tm)/ Lf σ = (ρcp) w /(ρcp) f
Th high constant temperature on the inner wall of the inner τv hydrodynamic dimensionless relaxation time of the fluid
cylinder (K ) τg fluid internal energy dimensionless relaxation time
Tc low constant temperature on the wall of the outer cylinder τgw wall internal energy dimensionless energy relaxation time
(K )
Tm melting temperature of the PCM (K )

During the last two decades, LBM has found its way towards si- made for the shape of the boundary [38].
mulating the melting phase change problems as well. The most popular To alleviate the above-mentioned problem, this paper, for the first
LBM method for simulating the melting phase change problems is the time, aims at using a body-fitted boundary layer-type non-uniform
enthalpy based method. This method was first introduced by Jiaung mesh with LBM to simulate the melting of a pure PCM confined within a
et al. [19] to simulate the conduction dominated phase change pro- cylindrical gap of an annulus with considering the effect of heat con-
blems. Chatterjee and Chakraborty [20] and Chakraborty and Chat- duction within the wall of the inner cylinder. The main reason behind
terjee [21,22] based on the enthalpy method developed a series of the using the body-fitted mesh is to be able to avoid the complicated curved
lattice Boltzmann models for conduction dominated and convection- boundary schemes and employ the boundary conditions introduced for
diffusion melting/solidification problems. Huber et al. [23] extended the straight boundaries, instead. In this study, the Taylor series least
the Jiaung et al.‘s [19] work for simulating the convection-diffusion square-based LBM (TLLBM) is employed to deal with the propagation
melting problems. The enthalpy method of Huber et al. [23], because of on the non-uniform mesh. Moreover, this paper aims at investigating
its simplicity, was used by so many researchers to numerically in- the effect of pertinent geometrical and thermophysical parameters as
vestigate the melting phase change problems [24–30]. Huang et al. [31] well as different thermal boundary conditions on the melting of the
introduced a new lattice Boltzmann method for modeling the phase PCM confined within a cylindrical annulus.
change problems based on the total enthalpy method. Those authors
modified the temperature equilibrium distribution function in such a 2. Macroscopic governing equations and the body-fitted mesh
way that the non-linear latent heat source in the energy equation dis-
appeared [31]. Many researchers successfully employed the Huang Fig. 1 shows a schematic diagram of the problem under investiga-
et al.'s [31] method to simulate different phase change problems tion in the present paper. As illustrated in Fig. 1, a horizontal concentric
[32–35]. cylindrical annulus is filled with a pure PCM initially at the melting
However, there are few works in the literature used the LBM method temperature Tm . The inner wall of the inner cylinder (r = rii ) is subject to
to simulate the PCM melting in the cylindrical LHTESS. Jourabian et al. a constant high temperature of Th (Th > Tm ) whereas, the wall of the
[36] used the Huber et al.‘s [23] method to investigate the nanoparticle outer cylinder (r = ro ) is considered to be at a constant low temperature
enhanced melting of a PCM in a cylindrical annulus. Luo et al. [37] of Tc (Tc = Tm ) or insulated. It should be emphasized that both mentioned
employed both Huber et al.'s [23] and Huang et al.'s [31] methods to boundary conditions for the outer cylinder are used by different re-
investigate the natural convection melting of a PCM in a shell con- searchers in numerical simulations, however, in most of the empirical
taining multiple tubes. It should be stated that all the above-mentioned researches the adiabatic boundary condition is used [39]. The outer
researchers used the standard LBM on a uniform Cartesian mesh. wall of the inner cylinder (r = rio ) is a conjugate interface where the
The lack of researches in which the LBM method is used to nu- thermal compatibility conditions should be satisfied between the two
merically simulate the melting of the PCM in cylindrical LHTESS, media namely, the wall of the inner cylinder and the PCM. The ther-
however, can be attributed to the difficulties encountered in applying mophysical properties of the PCM are presented in Table 1. It should be
the boundary conditions at the curved boundaries associated with the mentioned that in this study the same thermophysical properties are
geometries of those systems. That is, when the standard LBM with assumed for the solid and liquid PCM's [23–25].
uniform Cartesian mesh is used to simulate the PCM melting problem in The dimensionless forms of the continuity, momentum, energy
the cylindrical LHTESS, it is difficult to apply the hydrodynamic and conservation for the PCM and inner cylinder's wall media describing the
thermal boundary and interface conditions at the curved boundaries, physical problem of a two-dimensional natural convection-dominated
because in those cases the boundaries reside between the mesh points melting of a pure PCM are presented by Eqs. (1)–(4), respectively
which brings about errors to some extent because of the approximation [23,40].

2
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

made for the shape of the boundary. Moreover, with the body-fitted
meshes, one can use most of the LBM boundary condition schemes in-
troduced for the straight boundaries for the curved boundaries as well.

3. Methodology

3.1. The standard thermal lattice Boltzmann method

The 2D standard LBM with a body force uses the density and in-
ternal energy evolution equations to update, respectively, the density
and internal energy distribution functions exactly at the grid points on a
uniform Cartesian mesh, through two main steps called collision and
propagation as follows:

Collision:
1
f i∗ (r,t ) = fi (r,t ) − [f (r,t ) − fieq (r,t )] +
τv i
δtFi
1
gi∗ (r,t ) = gi (r,t ) − [g (r,t ) − gieq (r,t )]
τg i
Fig. 1. The schematic of the geometry and boundary conditions. (6)

Table 1 Propagation:
Thermophysical properties of the PCM. fi (r+ei δt , t + δt ) = f i∗ (r,t )
Property Value gi (r+ei δt , t + δt ) = gi∗ (r,t ) (7)
ρ (kg/m3 ) 997.1
where ei and ωi are, respectively, the particle velocity and weighting
μ (Pa s ) 8.9 × 10−4 factor in the ith lattice direction. fi (r,t ) and gi (r,t ) are, respectively, the
cp (J/kg K ) 4179
pre-collision density and internal energy distribution functions in the ith
k (W/m K ) 0.6
lattice direction at position r= (x , y ) and time t , f i∗ (r,t ) and gi∗ (r,t ) are,
β (1/ K ) 2.1 × 10−4
Pr 6.2 respectively, the post-collision density and internal energy distribution
Tm 0 ∘C functions in the ith lattice direction at position r= (x , y ) and time t . δt is
the time step, τv and τg are, respectively, the hydrodynamic and fluid
internal energy dimensionless relaxation times. fieq and gieq are the
It should be mentioned that in deriving Eqs. (1)–(4), it is assumed density and fluid internal energy equilibrium distribution functions,
that the melted PCM is a Newtonian fluid, the flow is transient and and Fi is the discrete body force in ith lattice direction. For the D2 Q9
remains in the laminar regime, and the Boussinesq approximation is lattice arrangement employed in this study (see Fig. 3), fieq , gieq , and Fi
invoked for the buoyancy term. Moreover, in deriving Eqs. (1)–(4), the are calculated, respectively, from Eqs. (8)–(10).
thermophysical properties of the solid and liquid PCMs are considered
e .V VV:ei ei − cs2 I ⎤
to be equal to each other and remain constant [23–25]. fieq = ωi ρf ⎡1 + i 2 +

⎣ cs 2cs4 ⎥
⎦ (8)
∇. V ∗ = 0 (1)

∂V ∗ e .V VV:ei ei − cs2 I ⎤
+ (V ∗. ∇) V ∗ = −∇p∗ + Pr∇2 V ∗ + RaPrΘf j gieq = ωi Tf ⎡1 + i 2 +
∂Fo (2) ⎢ c 2cs4 ⎥ (9)
⎣ s ⎦
∂Θf 1 ∂Fl
+ V ∗. ∇Θf = ∇2 Θf − 1 ⎞ ⎡ ei .F VF:ei ei V. F
∂Fo Ste ∂Fo (3) Fi = ωi ρf ⎛1 −

2
⎟+ − 2 ⎤
⎝ 2τv ⎠ ⎢
⎣ cs cs4 cs ⎥⎦ (10)
σ ∂Θ
⎛ ⎞ w = ∇2 Θw
⎝ λ ⎠ ∂Fo (4)
where, Fl is the local melt fraction. The other dimensionless parameters
and groups introduced to derive equations (1)–(4) are as follows:
(x , y ) VLc T − Tm αf t pLc2
(x ∗, y∗ ) = , V∗ = , Θ= , Fo = , p∗ = ,
Lc αf Th − Tm Lc2 ρf α 2f

gβ (Th − Tm) Lc3 (ρ cp) w kw cp (Th − Tm) ν


Ra = αf v
, σ= (ρ cp) f
, λ= kf
, Ste = Lf
, Pr = αf

(5)
where Pr , Ste , Fo and Ra are, respectively, the Prandtl, Stefan, Fourier,
and Rayleigh numbers, λ is the wall to PCM thermal conductivity ratio,
σ is the wall to PCM volumetric heat capacity ratio, Lc is a proper length
scale taken as 2(ro − rii ) in this study, g is the gravitational acceleration,
and β is the thermal expansion coefficient.
Fig. 2 demonstrates a typical body-fitted boundary layer-type non-
uniform mesh generated in this study for TLLBM simulations. As dis-
cussed before, when a body-fitted mesh is used, the geometry of the
curved boundaries can be modeled accurately because the mesh points Fig. 2. A typical body-fitted boundary layer-type non-uniform mesh used in this
reside exactly on the boundary, therefore, no approximation needs to be study.

3
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

between the enthalpies of the solid PCM, Ens and the melted PCM, Enl
as follows:

⎧0 Enn, k < Ens = cp Tm


⎪ Enn, k − Ens
Fln, k = Ens ≤ Enn, k ≤ Ens + Lf
⎨ Enl − Ens
⎪1 Enn, k > Ens + Lf
⎩ (16)

Finally, with knowing the local melt fractions at the current and
previous time steps, the collision of the internal energy distribution
function is modified for including the melting phase change as follows:

1 Lf
gi∗ (r, t ) = gi (r, t ) − [g (r, t ) − gieq (r, t )] − ωi [Fln, k (r) − Fln − 1, k (r)]
Fig. 3. The D2 Q9 lattice model. τg i cp
(17)
where cs is the lattice sound speed related to the lattice velocity c as This iteration procedure for a time step continues until the local
cs = c/ 3 , the lattice velocity c is equal to δx / δt , Tf is the fluid tem- melt fraction and the temperature fields converge to within a specified
perature, V is the macroscopic fluid velocity vector, and tolerance. However, over the selected ranges of parameters in this
F= ρf gβ (Tf − Tm) is the buoyancy force derived using the Boussinesq study, the iteration procedure is shown to have a negligible effect on
approximation, where Tm is the melting temperature of the pure PCM. the results as also mentioned by Huber et al. [23].
It should be mentioned that for the solid phase (i.e., the wall of the Noteworthy is the fact that Eq. (3) can be recovered exactly by
inner cylinder), a separate internal energy distribution function gwi is applying the Chapman-Enskog expansion to the modified evolution
used with an evolution equation similar to that of fluid (see Eqs. (6) and equation of the internal energy distribution function described in Eq.
(7)). The solid internal energy equilibrium distribution function also (17).
can be calculated by setting V= 0 in Eq. (9).
For the D2 Q9 lattice arrangement, the discrete velocity ei and the
weight factor ωi in each direction i are given by Eqs. (11) and (12), 3.3. Modifying the propagation step by TLLBM for the non-uniform meshes
respectively.
When the non-uniform mesh is employed in LBM, a problem arises
⎧ (0,0) i = 0 when it comes to performing the propagation process. Consider, after
ei = (cos θi , sin θi) c ; θi = (i − 1) π /2 i = 1,2,3,4 the propagation process takes place on a non-uniform mesh, say from a

typical node like A in a typical direction as shown in Fig. 4, the dis-
⎩ 2 (cos θi , sin θi ) c ; θi = (i − 5) π /2 + π /4 i = 5,6,7,8 (11)
tribution function may not coincide with the neighbor grid point in that
direction (i.e., point P in Fig. 4). As a result, it resides somewhere in
⎧ 4/9 i=0
ωi = 1/9 i = 1,2,3,4 between the grid points which in this case is shown by a point A′ in
⎨ Fig. 4. This is problematic because we are interested in knowing the
⎩1/36 i = 5,6,7,8 (12)
values of the macroscopic quantities like density, velocity, and tem-
The macroscopic quantities such as fluid density, velocity, internal perature exactly at the grid points not somewhere in between them. The
energy, and boundary heat flux can be calculated from the moments of TLLBM was proposed by Shu et al. [41] to resolve the above-mentioned
the density and internal energy distribution functions as follows: shortcoming of the standard LBM in dealing with the non-uniform
8 8 8 meshes. Shu et al. [42] theoretically proved that the TLLBM recovered
ρ = ∑i = 0 fi , V= ∑i = 0 fi ei / ρf + δt /(2ρf ) F, T = ∑i = 0 gi
the Navier-Stokes equation with the second order of accuracy, which is
τg − 0.5 8
q=(ρcp) ∑i = 0 gi ei , consistent with the accuracy of the standard LBM. In fact, the TLLBM
τg (13)
uses the Taylor series expansion in the spatial directions to find the
In the D2 Q9 model, from a Chapman-Enskog expansion, the fol-
lowing relations can be derived for the fluid kinematic viscosity ν , fluid
thermal diffusivity αf , and solid wall thermal diffusivity α w , respec-
tively, as presented by Eq. (14).

ν = cs2 δt (τv − 0.5), αf = cs2 δt (τg − 0.5), α w = cs2 δt (τgw − 0.5) (14)

3.2. Melting treatment in LBM

The enthalpy-based method of Huber et al. [23], which is an


iterative procedure to simultaneously find the melt fraction and tem-
perature, is employed in this study to model the natural convection-
dominated melting of a pure PCM filled in a cylindrical annulus. In this
method, a local enthalpy at time step n and iteration k is defined as
follows:

Enn, k = cp T n, k + Fln, k − 1 Lf (15)

where T n, k is the local temperature at time step n and iteration k , and


Fln, k − 1 is the local melt fraction at time step n and the previous iteration,
k − 1, respectively.
The next step is to find the melt fraction at time step n and iteration
k from a linear interpolation of the local enthalpy, available in Eq. (15), Fig. 4. A schematic of the propagation on a non-uniform mesh.

4
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

density and internal energy distribution functions at grid points at time- density populations at the 9 selected points at a time t in the direction i ,
level t + δt , to complete the propagation process. For example, in the known from the previous iteration. Noteworthy is the fact that matrix
case of the propagation of the density distribution function from the [A] i can be calculated prior to the solution for all time steps, only by
point A , as shown in Fig. 4, the second-order Taylor series expansion is knowing the coordinates of the grid points, time step, δt , and the lattice
written as Eq. (18). model. After determining the matrix[A] i , only its first row needs to be
stored to be used in the propagation process in all time steps. It should
∂fi (P , t + δt ) ∂fi (P , t + δt )
fi (A′ , t + δt ) = fi (P , t + δt ) + ΔxA + ΔyA be noted that this could reduce the computational costs.
∂x ∂y
The same procedure as discussed above can easily be repeated for
∂2fi (P , t + δt ) (ΔxA)2 ∂2fi (P , t + δt ) (ΔyA )2 the internal energy distribution function as well. However, in the pre-
+ 2
+
∂x 2 ∂y 2 2 sent study, because the same hydrodynamic and thermal lattice models
2
∂ fi (P , t + δt ) are chosen, there is no need to repeat Eqs. 18–24 for the internal energy
+ ΔxA ΔyA + O [(ΔxA)3, (ΔyA )3]
∂x ∂y (18) distribution function too. That is, in this case, the final coefficient
matrix [A] i obtained from Eq. (24) for the density distribution function
where ΔxA = (xA + eix δt ) − x p and ΔyA = (yA + eiy δt ) − yp . can also be used to find the unknown pre-collision internal energy
One can deduce from Eqs. (6), (7) and (13) that the left-hand side of distribution functions at the grid points as well, to complete the internal
Eq. (18) is actually the post-collision density distribution function at the energy propagation process as shown by Eq. (25).
point A at time-level t , known from the previous iteration. However,
8
the right-hand side of Eq. (18) includes six unknowns including one
distribution function, two first-order derivatives, and 3 s-order deriva-
V1i = gi (P , t + δt ) = ∑ a1,i k gik∗ , i=0−8
k=0 (25)
tives of the pre-collision density distribution function at point P and
time-level t + δt . To find these six unknowns appeared in Eq. (18), six where is the first row of the matrix [A] i and
a1,i k gik∗
is the post-collision
equations are needed. So, other than point A , Eq. (18) should be re- internal energy populations at the 9 selected points at a time t , known
peated for five other points, which can be chosen arbitrarily from the from the previous iteration. To study further about the TLLBM, readers
nearby points. Shu et al. [43] selected these points as P , B , C , D , and H are referred to [41,43–47].
(see Fig. 4).
The six above-mentioned equations can be written in a matrix form 4. Implementation of the boundary and interface conditions in
for all the lattice directions i = 0 − 8 (see Fig. 4) as follows: LBM
{f }∗i = [S ]i {V }i i=0−8 (19)
As seen in Fig. 1, the geometry of the problem under investigation is
In Eq. (19), [S ]i is a 6 × 6 coefficient matrix, {V }i is the unknown a cylindrical annulus filled with a pure PCM which is initially at its
vector, and {f }∗i is the known vector. The known vector {f }∗i , coefficient melting temperature Tm . As illustrated in Fig. 1, this geometry has three
matrix[S ]i , and the unknown vector {V }i are presented, respectively, by boundaries namely, the inner wall of the inner cylinder (r = rii ), the
Eqs. (20)–(22). outer wall of the inner cylinder (r = rio ), which is a conjugate interface,
{f }∗i = [f i∗ (P , t ), f i∗ (A, t ), f i∗ (B, t ), f i∗ (C , t ), f i∗ (D , t ), f i∗ (H , t )]T , i and the wall of the outer cylinder (r = ro ). The no-slip condition is
applied to all three above-mentioned boundaries. At r = rii a constant
=0−8 (20)
higher temperature of Th (Th > Tm ) is applied whereas, at r = ro a con-
stant lower temperature of Tc (Tc = Tm ) or the adiabatic condition is
(ΔxP )i (ΔyP )i (ΔxP )i2 /2 (ΔyP )i2 /2 (ΔxP )i (ΔyP )i ⎤
⎡1 considered. Finally, at the conjugate interface, r = rio the compatibility
(ΔxA)i (ΔyA )i (ΔxA)i2 /2 (ΔyA )i2 /2 (ΔxA)i (ΔyA )i ⎥
⎢1
⎢ ⎥ conditions (i.e., continuity of temperatures and normal heat fluxes) are
(ΔxB )i (ΔyB )i (ΔxB )i2 /2 (ΔyB )i2 /2 (ΔxB )i (ΔyB )i ⎥
⎢1 considered.
[S ]i = ⎢ ⎥, =i
(Δx C )i (ΔyC )i (Δx C )i2 /2 (ΔyC )i2 /2 (Δx C )i (ΔyC )i ⎥
⎢1 All the above-mentioned boundary and conjugate interface condi-
⎢ ⎥ tions are given in terms of the macroscopic quantities such as velocity
(ΔxD )i (ΔyD )i (ΔxD )i2 /2 (ΔyD )i2 /2 (ΔxD )i (ΔyD )i ⎥
⎢1 and temperature, or their gradients. Therefore, these macroscopic
(ΔxH )i (ΔyH )i (ΔxH )i /2 (ΔyH )i /2 (ΔxH )i (ΔyH )i ⎥
⎢1 2 2
⎣ ⎦ boundary conditions need to be transformed into the LBM framework.
=0−8 (21) In the present research, the extrapolation method of Guo et al. [48] is
employed for all the no-slip boundary conditions and the counter-slip
{V }i = (fi , ∂fi / ∂x , ∂fi / ∂y, ∂2fi / ∂x 2 , ∂2fi / ∂y 2 , ∂2f / ∂x ∂y )T , i=0−8 (22) internal energy method of D'Orazio et al. [49] is used for the constant
temperature or adiabatic boundary conditions. Care should be taken
It should be noted that we are only interested in finding the first
that the method of D'Orazio et al. [49] was originally proposed for the
element of the vector {V }i , which is V1i = fi (P , t + δt ) .
straight boundaries and cannot be used for curved boundaries in the
In practical applications, the matrix [S ]i might be singular or ill-
case of uniform Cartesian meshes. However, because in this study the
conditioned [43]. To fix this problem, Shu et al. [43–45] suggested to
non-uniform body-fitted mesh is used, the method of D'Orazio et al.
use more than six points (i.e., nine points in 2D simulations as shown in
[49] can be used for the curved boundaries as well. In this study, the
Fig. 4) and optimize this approximation by applying the least square
compatibility conditions at the wall-PCM conjugate interface are also
method to Eq. (19) to get the solution of the unknown vector {V }i as Eq.
employed using the extrapolation method of Guo et al. [48].
(23).
To save some space, only the conjugate heat transfer formulation at
−1
{V }i = ([S ]Ti [S ]i ) [S ]Ti {f }∗i = [A]i {f }∗i , i=0−8 (23) the conjugate interface is discussed here. Consider a typical node on the
wall-PCM conjugate interface denoted by “Int” in Fig. 5. At this node,
Those above-mentioned additional three points are taken as F , G ,
there are some unknown internal energy populations entering the PCM
and E as shown in Fig. 4. Now, the size of the coefficient matrix [S ]i will
medium from the wall and vice versa. Guo et al. [48] proposed to de-
be 9 × 6. From Eq. (23), the first element of the unknown vector {V }i ,
compose those unknown populations to their equilibrium and non-
we are interested in, can be presented by Eq. (24).
equilibrium parts as gi (r,t )= gieq (r,t )+ gineq (r,t ) . The equilibrium part,
8 then, can be calculated from the knowledge of the temperature at the
V1i = fi (P , t + δt ) = ∑ a1,i k fik∗ , i=0−8 interface node, whereas, the non-equilibrium part is extrapolated from
k=0 (24)
the neighboring points.
where a1,i k is the first row of the matrix [A] i and f ik∗ is the post-collision To calculate the equilibrium part, first the interface temperature is

5
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

dimensionless temperature distribution along the radial direction at


θ = 30∘ in the present study (equivalent to θ = 120∘ in the work of
Kuhen and Goldstein [50]) with the experimental data of Kuhen and
Goldstein [50]. As seen in Fig. 6, the result of the present study is in a
good agreement with the experimental data of Kuhen and Goldstein
[50].
Next, the ability of the TLLBM FORTRAN code for simulating the
melting phase change problems is evaluated. To do so, the natural
convection melting of a phase change material confined in a square
enclosure is considered with Ra = 5 × 10 4 , Pr = 1, and Ste = 10 . Fig. 7
compares the melt fraction versus the modified dimensionless time
FoSte of the present research with that of Huber et al. [23]. As seen in
Fig. 7, the results of the present study is in an excellent agreement with
that of Huber et al. [23].

5.2. Mesh independency study


Fig. 5. A schematic of the conjugate interface.
To ensure that the final results are mesh independent, melting of the
calculated by satisfying the compatibility conditions (i.e., continuity of pure PCM in a cylindrical annulus without considering the thickness of
temperature and normal heat flux) at the conjugate interface as follows: the inner cylinder is simulated for Ra = 105 and Ste = 1. Six non-uni-
form meshes of 61 × 168, 81 × 208, 101 × 248, 121 × 296, 145 × 352 , and
ΔrPCM
TPN + λ T
Δrw WN 173 × 418 are considered, in which, the first and second numbers denote
TInt = Δr the number of mesh points in the radial and tangential directions, re-
1+ λ ΔrPCM (26)
w spectively. Fig. 8 shows the variation of the liquid fraction versus the
where λ is the wall to PCM thermal conductivity ratio, TInt is the in- FoSte for the different mesh sizes. It is seen that the results of the me-
terface temperature at a typical node, TWN and TPN are the temperatures shes 145 × 352 and 173 × 418 are within less than 0.8%, therefore, the
of the neighboring nodes within the wall and the PCM media, respec- mesh 145 × 352 is selected for the rest of this paper in favor of reducing
tively, ΔrPCM and Δrw are, respectively, the radial distances from the the computational cost of the numerical simulations yet keeping the
interface to the next row of the mesh points within the PCM and the accuracy in an acceptable range for the engineering applications.
wall media.
The non-equilibrium part of the unknown internal energy popula- 6. Results and discussions
tion is taken equal to that of the first neighbor point in the normal
direction, which in the case of the PCM, for example, is point PN (see According to Eqs. (1)–(4), six dimensionless numbers and groups
Fig. 5), gineq (r,t )= [(gi (r+ΔrPCM n, t ) − gieq (r+ΔrPCM n, t ))]. It should be describe the behavior of the natural convection-dominated melting of a
mentioned that even though the first-order extrapolation is used for PCM encapsulated in a cylindrical annulus with conjugate heat transfer
determining the non-equilibrium part, Guo et al. [48] showed that this within the wall of the inner cylinder. These dimensionless numbers and
method was of the second order overall accuracy. groups are the Rayleigh number Ra, Prandtl number Pr , Stefan number
Once the equilibrium and non-equilibrium parts of the unknown Ste , Fourier number Fo , wall to PCM thermal conductivity ratio λ , and
internal energy populations at the conjugate interface are known, the wall to PCM volumetric heat capacity ratio σ . The effect of the geo-
post-collision populations for the wall and PCM media are calculated as metry of the cylindrical annulus is also included by introducing a
follows:

1
gi∗ (r, t ) = ωi TInt + ⎜⎛1 − ⎟⎞ [(gi (r+ΔrPCM n, t ) − gieq (r+ΔrPCM n, t ))]
⎝ τg⎠

(27)

1 ⎞

gwi (r, t ) = ωi TInt + ⎛⎜1 − eq
⎟ [(gwi (r− Δrw n, t ) − g wi (r− Δrw n, t ))]
⎝ τgw ⎠

(28)
where, TInt is the temperature of a typical node on the wall-PCM con-
jugate interface, and n is the unit vector normal to the wall-PCM con-
jugate interface (see Fig. 5).

5. Verification of the TLLBM FORTRAN code and mesh


independence analysis

5.1. Verification of the TLLBM FORTRAN code

In this section, at first, the developed TLLBM FORTRAN code for the
natural convection of air in a concentric cylindrical annulus with con-
jugate heat transfer within the wall of the inner cylinder is compared
with the experimental data of Kuhen and Goldstein [50] for
Ra = 4.7 × 10 4 and Pr = 0.706 . The other geometrical and thermo-
physical parameters used in this test case are selected as presented in Fig. 6. Comparison of the radial temperature distribution with the experi-
Kuhen and Goldstein [50]. Fig. 6 compares the result of the mental data of the Kuhen and Goldstein [50] at.θ = 30∘

6
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

of the inner cylinder is not taken into account. The Rayleigh, Prandtl,
and Stefan numbers for this case are considered to be Ra = 105 ,
Pr = 6.2 , and Ste = 1, with the radius ratio equal to rr = 2.25. As seen in
Fig. 9, at the early stages of the melting process the dimensionless
isotherms within the cylindrical annulus are annular and parallel to
each other, indicating the dominance of the conduction heat transfer
mechanism in those early times. As the time goes on further and fur-
ther, the convection heat transfer mechanism becomes more dominant
and two symmetrical circulation zones appear within the left and right
halves of the annulus. That is, as seen in Fig. 9b, c, and d, where, a hot
plume of the melted PCM develops from the upper part of the inner
cylinder and deforms the annular form of the isotherms within the
upper part of the annulus, implying the dominance of the convection
mechanism in that region. It should be noted that this is not the case for
the lower part of the cylindrical annulus where the conduction me-
chanism is dominant during the whole time of the melting process.
From Fig. 9, one can also observe that around FoSte = 0.065 the melted
PCM touches the wall of the outer cylinder for the first time.
Fig. 10 is plotted to show the evolution of the liquid fraction defined
as the volume of the liquid region to the volume of the gap of the cy-
lindrical annulus. As seen in Fig. 10, the fraction of the PCM melted
within the cylindrical annulus increases as the dimensionless time in-
Fig. 7. Comparison of the evolution of the liquid fraction in a square enclosure creases, however, the melting rate changes during the whole process.
with the work of Huber et al. [23]. To more elaborate on this, three distinct time zones are identified for
the melting of the PCM in a cylindrical annulus, denoted as zones A, B,
and C as shown in Fig. 10. The zone A contains the time interval during
which the conduction heat transfer mechanism is dominant as discussed
in Fig. 9 as well. At the early times of the zone A, the melting rate is
more pronounced because the annular melted region (see Fig. 9a) de-
veloped around the inner cylinder at the early times of the melting
process, has a small thickness, thereby causing a small conduction
thermal resistance to the heat transferred from the inner cylinder to the
PCM. As seen in Fig. 10, at the later times in the zone A, the melting rate
decreases as the thickness of the annular melted region increases,
thereby applying a bigger conduction thermal resistance to the heat
flow from the inner cylinder to the PCM. However, at the same time, the
buoyancy force in the melted region promotes, thereby increasing the
share of the convection heat transfer mechanism. As a result, the de-
crease in the melting rate stops and it remains constant throughout the
zone B until the melted PCM touches the wall of the outer cylinder for
the first time (see Fig. 9c). After that, in the zone C, the melting rate
decreases until the end of the melting period. The same result, as dis-
cussed above, has been reported by Huber et al. [23] for the melting of
a pure PCM confined in a square enclosure.

6.1. Effect of the Rayleigh number

Fig. 11 shows the effect of the Rayleigh number on the time evo-
Fig. 8. Mesh independency test of the results of the TLLBM FORTRAN code
developed to simulate melting of a PCM confined in a cylindrical annulus. lution of the liquid fraction in a cylindrical annulus with rr = 2.25
without considering the wall thickness for the inner cylinder. As seen in
Fig. 11, the higher Rayleigh numbers result in the higher liquid fraction
geometrical parameter called the radius ratio rr defined as rr = ro/ rii for the same FoSte . However, as shown in Fig. 11, the effect of Rayleigh
(see Fig. 1). The values of the above-mentioned non-dimensional number on the liquid fraction is more pronounced in zone B (refer to
numbers and groups plus the non-dimensional inner wall thickness Fig. 10) which is the convection dominated zone, whereas, for the zones
t ∗ = t /[2(ro − rii )] considered in this research are presented in Table 2.
As discussed before in Sec. 2, the constant temperature boundary con- Table 2
ditions are applied to the inner wall of the smaller cylinder and the wall Simulation parameters.
of the outer cylinder for all the presented results, unless otherwise is
Parameters Value
mentioned. Moreover, in all the cases presented in this paper, the pure
PCM is considered to be initially at the melting temperature and no Ra 10 4 , 5 × 10 4 , 105
subcooling is considered. Pr 6.2
Fig. 9 shows the dimensionless isotherms, local melt fraction, and Ste 1
streamlines of the pure PCM melting problem for different values of the λ 1,10,100
σ 1
modified dimensionless time FoSte . The pure PCM is considered to be
t∗ 0.025, 0.05,0.1
initially at the melting temperature of Tm = 0∘C and is confined in a rr 2.25, 2.12, 2, 1.8
cylindrical annulus. It should be stated that at this stage, the thickness

7
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

Fig. 9. Dimensionless isotherms, local melt fraction, and the streamlines for melting of the PCM in a cylindrical annulus with rr = 2.25, Ra = 105 , without considering
the wall thickness for the inner cylinder (a) FoSte = 0.01 (b) FoSte = 0.04 (c) FoSte = 0.065 (d).FoSte = 0.119

Fig. 11. Effect of the Rayleigh number on the evolution of the liquid fraction in
Fig. 10. Evolution of the liquid fraction for melting of PCM confined within a the cylindrical annulus without considering the wall thickness.
cylindrical annulus with rr = 2.25 and.Ra = 105

6.2. Effect of the cylinders radius ratio


A and C (see Fig. 10) which are, respectively, the conduction dominated
zone and the zone in which the melted PCM is in touch with the Fig. 12 demonstrates the evolution of the liquid fraction for dif-
boundary, the effect of Rayleigh number is negligible for the range of ferent values of the radius ratio rr = 1.8, 2, 2.12, 2.2 . To generate these
Rayleigh numbers studied here. different values of the radius ratio, a constant radius is considered for
the outer cylinder and the radius of the inner cylinder is increased in

8
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

before, the liquid fraction is defined as the volume of the liquid region
to the volume of the gap of the cylindrical annulus. Because the volume
of the gap decreases by decreasing the radius ratio, from Fig. 12, one
can not conclude that for the smaller radius ratios the volume of the
melted region is absolutely greater than that of the bigger ratios.
However, in this study, it is shown that the volume of the melted region
also increases by decreasing the radius ratio.

6.3. Effect of the thermal conductivity ratio

In this section, the effect of the conducting wall of the inner cylinder
on the evolution of the dimensionless isotherms is investigated for three
wall to PCM conductivity ratios λ = 1, 10, 100 as demonstrated by
Fig. 13. Fig. 13, shows the evolution of the dimensionless isotherms for
the melting of the pure PCM within a cylindrical annulus with the ra-
dius ratio of rr = 2.25 and the dimensionless wall thickness of t ∗ = 0.05.
As the first result from Fig. 13, it is seen that the conduction heat
transfer is the dominant heat transfer mechanism at the early stages of
the melting process whereas, at the later times, the convection heat
transfer plays a significant role in transferring heat from the inner cy-
linder to the pure PCM. The only difference between these results and
those of the case where the thickness of the wall of the inner cylinder is
Fig. 12. Effect of the radius ratio of the cylindrical annulus on the evolution of
neglected (see Fig. 9) is that for the low thermal conductivity ratio
the liquid fraction without considering the wall thickness.
(λ = 1) the dominance of the convection heat transfer is delayed com-
pared to the cases of the higher thermal conductivity ratios
order to generate the geometries with smaller radius ratios. As seen in (λ = 10, 100 ), which, in turn, slows down the melting process. Fur-
Fig. 12, the general trend of the evolution of the liquid fraction is the thermore, in Fig. 13, one can conclude that the results of evolution of
same for the four of the examined radius ratios. That is, the three dis- the isotherms for λ = 10 and λ = 100 are almost indistinguishable. This
tinct zones discussed in Fig. 10 can be seen here for all the radius ratios is a very important result and can be used to reduce the costs associated
as well. As the second result, one can see in Fig. 12 that changing the with the pipe material in the design of LHTESS of the cylindrical an-
radius ratio significantly affects the liquid fraction in a way that the nulus type. However, this result needs to be further investigated ex-
liquid fraction is bigger for the smaller radius ratios. As discussed perimentally.

Fig. 13. The evolution of dimensionless isotherms with rr = 2.25 and t ∗ = 0.05 for different thermal conductivity ratios λ (a) FoSte = 0.011 (b) FoSte = 0.04 (c)
FoSte = 0.069 (d).FoSte = 0.119

9
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

To more elaborate on the above-mentioned results, Fig. 14a and b


are plotted. Fig. 14a and b compare the evolution of the liquid fraction
within a cylindrical annulus for different thermal conductivity ratios
λ = 1, 10, 100 , against each other and against the corresponding cases
without considering the wall thickness for the inner cylinder, respec-
tively, for the dimensionless wall thicknesses t ∗ = 0.05 and t ∗ = 0.1. As
seen from Fig. 14a and b, for the examined values of λ , it is seen that the
results of λ = 10 and λ = 100 are very close to each other, however,
they significantly differ from that of λ = 1 for almost the entire range of
the modified dimensionless time FoSte . It is also noted in Fig. 14a and b
that the deviation of the results of λ = 1 from those of λ = 10 and
λ = 100 is more pronounced for the case with the bigger dimensionless
wall thickness. As another important result from Fig. 14a and b, one can
observe that the results of the evolution of the liquid fraction for the
high thermal conductivities λ = 10 and λ = 100 are in good agreement
with those of the cylindrical annulus without the wall thickness, in such
a way that for λ = 100 the difference between the two are almost in-
distinguishable. This is because, for the higher thermal conductivity
ratios, the temperature distribution within the wall of the inner cylinder
becomes almost uniform, resembling the constant temperature
boundary condition at the conjugate interface r = rio (see Fig. 1).

6.4. Effect of the wall thickness Fig. 15. The effect of the inner cylinder wall thickness on the evolution of the
liquid fraction for different λ .
Fig. 15 illustrates the effect of the dimensionless wall thickness on
the evolution of the liquid fraction for two thermal conductivity ratios Figs. 16–13, one can conclude that before the hot plume of the melted
namely, λ = 1 and λ = 100 , as representatives of the low and high PCM touches the wall of the outer cylinder, the shape of the di-
thermal conductivities, respectively. It is concluded from Fig. 15 that mensionless isotherms is similar for the cases with adiabatic and con-
the effect of the wall thickness on the liquid fraction for λ = 1 is com- stant temperature conditions. However, as seen in Fig. 16, after the
pletely against that of λ = 100 . That is, for λ = 1increasing the di- plume touches the wall of the outer cylinder, the shape of the isotherms
mensionless wall thickness decreases the liquid fraction whereas, for starts to change adjusting themselves to the adiabatic condition applied
λ = 100 the reverse is true. to the wall of the outer cylinder. As a result, because the heat can not
escape the boundary this enhances the melting rate at the final stage of
6.5. Effect of the boundary condition on the outer cylinder the melting process.
Fig. 18 compares the evolution of the liquid fraction for the cases
In this section, the adiabatic boundary condition is applied to the with constant temperature and adiabatic conditions on the wall of the
wall of the outer cylinder instead of the constant temperature condi- outer cylinder, for different thermal conductivity ratios. As illustrated
tion. Figs. 16 and 17, respectively, demonstrate the evolution of the in Fig. 18, for all the values of the examined thermal conductivity ratio,
dimensionless isotherms and local melt fraction corresponding to the the liquid fractions of the cases of the adiabatic condition are ap-
melting of the pure PCM confined within a cylindrical annulus with an proximately comparable to those of the constant temperature condition,
adiabatic condition applied to the wall of the outer cylinder. Comparing for the times before the liquid PCM first touches the wall of the outer

Fig. 14. The evolution of the liquid fraction for the different thermal conductivity ratios compared to the case without the wall thickness (a) t ∗ = 0.05 (b).t ∗ = 0.1

10
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

Fig. 16. The evolution of the dimensionless isotherms for the case of the adiabatic boundary condition on the wall of the outer cylinder for different conductivity
ratios (a) FoSte = 0.011 (b) FoSte = 0.04 (c) FoSte = 0.069 (d).FoSte = 0.119

cylinder, with the results of the adiabatic condition for the liquid 7. Conclusions
fraction being slightly bigger than those of the constant temperature
condition. After the liquid touches the outer cylinder, the adiabatic The body-fitted boundary-layer type non-uniform mesh is success-
condition on the wall accelerates the melting of the remained PCM, fully used with the Taylor series expansion least square based lattice
causing a bigger deviation between the results of the two boundary Boltzmann method (TLLBM) to simulate the melting of a pure PCM
conditions. That is, in case of the constant temperature boundary con- confined within the cylindrical annulus with conjugate heat transfer in
dition there is a heat leak from the outer cylinder which is responsible the inner wall. Because of the body-fitted mesh employed here, the
for decreasing the melting rate compared to the adiabatic boundary complicated LBM curved boundary and interface schemes are avoided,
condition. instead, the simpler schemes, established for the straight boundaries,
are used. However, comparing the computational costs of the TLLBM

Fig. 17. The evolution of the local melt fraction for the case with the adiabatic boundary condition on the wall of the outer cylinder (a) FoSte = 0.011 (b) FoSte = 0.04
(c) FoSte = 0.069 (d).FoSte = 0.119

11
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

[8] K.W. Ng, Z.X. Gong, A.S. Mujumdar, Heat transfer in free convection-dominated
melting of a phase change material in a horizontal annulus, Int. Commun. Heat
Mass Tran. 25 (5) (1998) 631–640.
[9] G.H. Bagheri, M.A. Mehrabian, K. Hooman, Numerical study of the transient be-
haviour of a thermal storage module containing phase-change material, Proc. IME J.
Power Energy 224 (4) (2010) 505–516.
[10] A.R. Darzi, M. Farhadi, K. Sedighi, Numerical study of melting inside concentric and
eccentric horizontal annulus, Appl. Math. Model. 36 (9) (2012) 4080–4086.
[11] M. Lacroix, Numerical simulation of a shell-and-tube latent heat thermal energy
storage unit, Sol. Energy 50 (4) (1993) 357–367.
[12] S. Addakiri, et al., Numerical study of melting/solidification by a hybrid method
coupling a lattice Boltzmann and a finite volumes approaches, in: A. Ochsner,
G.E. Murch, A. Shokuhfar (Eds.), Diffusion in Solids and Liquids Iv, Trans Tech
Publications Ltd, Stafa-Zurich, 2009, pp. 477–482.
[13] M. Jourabian, M. Farhadi, K. Sedighi, On the expedited melting of phase change
material (PCM) through dispersion of nanoparticles in the thermal storage unit,
Comput. Math. Appl. 67 (7) (2014) 1358–1372.
[14] S. Paria, et al., Performance evaluation of latent heat energy storage in horizontal
shell-and-finned tube for solar application, J. Therm. Anal. Calorim. 123 (2) (2016)
1371–1381.
[15] J. Liu, et al., Numerical investigation on the heat transfer enhancement of a latent
heat thermal energy storage system with bundled tube structures, Appl. Therm.
Eng. 112 (2017) 820–831.
[16] N.S. Dhaidan, J.M. Khodadadi, Melting and convection of phase change materials in
different shape containers: a review, Renew. Sustain. Energy Rev. 43 (Supplement
C) (2015) 449–477.
[17] N.S. Dhaidan, J.M. Khodadadi, Improved performance of latent heat energy storage
systems utilizing high thermal conductivity fins: a review, J. Renew. Sustain.
Energy 9 (3) (2017).
Fig. 18. Comparing the evolution of the liquid fractions of the cases with [18] X. He, S. Chen, G.D. Doolen, A novel thermal model for the lattice Boltzmann
constant temperature and adiabatic conditions for different conductivity ratios. method in incompressible limit, J. Comput. Phys. 146 (1) (1998) 282–300.
[19] W.S. Jiaung, J.R. Ho, C.P. Kuo, Lattice Boltzmann method for the heat conduction
problem with phase change, Numer. Heat Tran. Part B-Fundamentals 39 (2) (2001)
with the standard LBM employing the common curved boundary 167–187.
schemes is yet to be investigated, especially for more complicated [20] D. Chatterjee, S. Chakraborty, An enthalpy-based lattice Boltzmann model for dif-
fusion dominated solid-liquid phase transformation, Physics Lett. Section A: Gen.
geometries. It should be mentioned that the presented method is in-
Atom. Solid State Phys. 341 (1–4) (2005) 320–330.
dependent of the melting model and can be used with available im- [21] S. Chakraborty, D. Chatterjee, An enthalpy-based hybrid lattice-Boltzmann method
proved enthalpy methods as well. for modelling solid-liquid phase transition in the presence of convective transport,
In this paper, effects of pertinent geometrical (i.e., radius ratio rr J. Fluid Mech. 592 (2007) 155–175.
[22] D. Chatterjee, S. Chakraborty, An enthalpy-source based lattice Boltzmann model
and dimensionless inner wall thickness t ∗), thermophysical (wall to for conduction dominated phase change of pure substances, Int. J. Therm. Sci. 47
PCM conductivity ratio λ ), and flow (Rayleigh number) parameters, as (5) (2008) 552–559.
well as the outer cylinder thermal boundary condition, are investigated [23] C. Huber, et al., Lattice Boltzmann model for melting with natural convection, Int.
J. Heat Fluid Flow 29 (5) (2008) 1469–1480.
on the evolution of the liquid fraction. It is concluded from the results [24] M. Jourabian, M. Farhadi, A.A.R. Darzi, Simulation of natural convection melting in
that using a smaller radius ratio for the cylindrical annulus enhances an inclined cavity using lattice Boltzmann method, Sci. Iran. 19 (4) (2012)
the melting process. It is also figured out that for the inner cylinder, 1066–1073.
[25] M. Jourabian, M. Farhadi, A.A.R. Darzi, Lattice Boltzmann investigation for en-
using a material with λ = 100 gives the results which are in a good hancing the thermal conductivity of ice using Al2O3 porous matrix, Int. J. Comput.
agreement with those of λ = 10 . As an interesting result, it is seen that Fluid Dynam. 26 (9–10) (2012) 451–462.
increasing the dimensionless inner wall thickness enhances the melting [26] M. Jourabian, et al., Simulation of natural convection melting in a cavity with fin
using lattice Boltzmann method, Int. J. Numer. Methods Fluid. 70 (3) (2012)
process for the higher thermal conductivity ratios (λ = 10 and λ = 100 )
313–325.
whereas, for the lower thermal conductivity ratio (λ = 1) the reverse is [27] A.A.R. Darzi, M. Farhadi, M. Jourabian, Lattice Boltzmann simulation of heat transfer
correct. Finally, it is seen that the effect of applying the adiabatic enhancement during melting by using nanoparticles Iranian, Journal of Sci. Technol.-
Trans. Mech. Eng. 37 (M1) (2013) 23–37.
boundary condition on the outer cylinder, instead of the constant
[28] W. Zhu, M. Wang, H. Chen, 2D and 3D lattice Boltzmann simulation for natural
temperature, is to enhance the melting during the entire process, convection melting, Int. J. Therm. Sci. 117 (2017) 239–250.
however, this improvement is more pronounced for the times after the [29] R.K. Dai, et al., Evolution of natural convection melting inside cavity heated from
melted PCM first touches the outer cylinder. different sides using enthalpy based lattice Boltzmann method, Int. J. Heat Mass
Tran. 121 (2018) 715–725.
[30] M. Jourabian, M. Farhadi, A.R. Darzi, Constrained ice melting around one cylinder
Declarations of interest in horizontal cavity accelerated using three heat transfer enhancement techniques,
Int. J. Therm. Sci. 125 (2018) 231–247.
[31] R.Z. Huang, H.Y. Wu, P. Cheng, A new lattice Boltzmann model for solid-liquid
None to be declared. phase change, Int. J. Heat Mass Tran. 59 (2013) 295–301.
[32] D. Gao, et al., Lattice Boltzmann modeling of melting of phase change materials in
References porous media with conducting fins, Appl. Therm. Eng. 118 (2017) 315–327.
[33] D. Gao, et al., An improved lattice Boltzmann method for solid-liquid phase change
in porous media under local thermal non-equilibrium conditions, Int. J. Heat Mass
[1] M. Kenisarin, K. Mahkamov, Solar energy storage using phase change materials, Tran. 110 (2017) 58–62.
Renew. Sustain. Energy Rev. 11 (9) (2007) 1913–1965. [34] Y. Hu, et al., Lattice Boltzmann simulation for three-dimensional natural convection
[2] A. Sharma, et al., Review on thermal energy storage with phase change materials with solid-liquid phase change, Int. J. Heat Mass Tran. 113 (2017) 1168–1178.
and applications, Renew. Sustain. Energy Rev. 13 (2) (2009) 318–345. [35] D. Li, et al., Three-dimensional lattice Boltzmann models for solid-liquid phase
[3] F.L. Tan, C.P. Tso, Cooling of mobile electronic devices using phase change mate- change, Int. J. Heat Mass Tran. 115 (2017) 1334–1347.
rials, Appl. Therm. Eng. 24 (2) (2004) 159–169. [36] M. Jourabian, et al., Melting of NEPCM within a cylindrical tube: numerical study
[4] D. Zhou, C.Y. Zhao, Y. Tian, Review on thermal energy storage with phase change using the lattice Boltzmann method, Numer. Heat Tran. Part A: Applications 61 (12)
materials (PCMs) in building applications, Appl. Energy 92 (2012) 593–605. (2012) 929–948.
[5] A.A. Al-abidi, et al., CFD applications for latent heat thermal energy storage: a [37] K. Luo, et al., Lattice Boltzmann simulation of convection melting in complex heat
review, Renew. Sustain. Energy Rev. 20 (Supplement C) (2013) 353–363. storage systems filled with phase change materials, Appl. Therm. Eng. 86 (2015)
[6] F. Agyenim, et al., A review of materials, heat transfer and phase change problem 238–250.
formulation for latent heat thermal energy storage systems (LHTESS), Renew. [38] Z. Guo, S. Chang, Lattice Boltzmann Method And its Applications In Engineering
Sustain. Energy Rev. 14 (2) (2010) 615–628. Advances in Computational Fluid Dynamics vol. 3, World Scientific, 2013.
[7] C.J. Ho, K.C. Lin, Numerical simulation of melting inside a cylindrical annulus, [39] L. Fan, J.M. Khodadadi, Thermal conductivity enhancement of phase change ma-
Model. Simulat. Contr. B 3 (3) (1985) 25–48. terials for thermal energy storage: a review, Renew. Sustain. Energy Rev. 15 (1)

12
G. Imani International Journal of Thermal Sciences xxx (xxxx) xxx–xxx

(2011) 24–46. method: an efficient approach for simulation of incompressible viscous flows, Prog.
[40] E.A. Semma, M. El Ganaoui, R. Bennacer, Lattice Boltzmann method for melting/ Comput. Fluid Dynam. Int. J. 5 (1–2) (2005) 27–36.
solidification problems, Compt. Rendus Mec. 335 (5–6) (2007) 295–303. [46] C. Shu, et al., Application of Taylor series expansion and Least-squares-based lattice
[41] C. Shu, X.D. Niu, Y.T. Chew, Taylor-series expansion and least-squares-based lattice Boltzmann method to simulate turbulent flows, J. Turbul. 7 (38) (2006) 1–12.
Boltzmann method: two-dimensional formulation and its applications, Phys. Rev. [47] C. Shu, et al., Numerical simulation of flows past a rotational circular cylinder by
65 (3) (2002). Taylor-series-expansion and least squares-based lattice Boltzmann method, Int. J.
[42] C. Shu, Y.T. Chew, X.D. Niu, Least-squares-based lattice Boltzmann method: a Mod. Phys. C 16 (11) (2005) 1753–1770.
meshless approach for simulation of flows with complex geometry, Phys. Rev. 64 [48] Z. Guo, C. Zheng, B. Shi, An extrapolation method for boundary conditions in lattice
(4) (2001) 045701. Boltzmann method, Phys. Fluids 14 (6) (2002) 2007–2010.
[43] C. Shu, Y. Peng, Y.T. Chew, Simulation of natural convection in a square cavity by [49] A. D'Orazio, S. Succi, Boundary conditions for thermal lattice Boltzmann simula-
Taylor series expansion- and least squares-based lattice Boltzmann method, Int. J. tions, in: M.A.P. Sloot, et al. (Ed.), Computational Science - Iccs 2003, Pt I,
Mod. Phys. C 13 (10) (2002) 1399–1414. Proceedings, Springer-Verlag Berlin, Berlin, 2003, pp. 977–986.
[44] C. Shu, X.D. Niu, Y.T. Chew, Taylor series expansion and least squares-based lattice [50] T.H. Kuehn, R.J. Goldstein, An experimental and theoretical study of natural con-
Boltzmann method: three-dimensional formulation and its applications, Int. J. Mod. vection in the annulus between horizontal concentric cylinders, J. Fluid Mech. 74
Phys. C 14 (7) (2003) 925–944. (4) (1976) 695–719.
[45] C. Shu, et al., Taylor series expansion- and least square-based Lattice Boltzmann

13

Potrebbero piacerti anche