Está en la página 1de 9

International Journal of Heat and Mass Transfer 55 (2012) 2420–2428

Contents lists available at SciVerse ScienceDirect

International Journal of Heat and Mass Transfer


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

An implicit lattice Boltzmann model for heat conduction with phase change
Mohsen Eshraghi, Sergio D. Felicelli ⇑
Mechanical Engineering Department and Center for Advanced Vehicular Systems, Mississippi State University, MS 39762, USA

a r t i c l e i n f o a b s t r a c t

Article history: In this work, a new variation of the lattice Boltzmann method (LBM) was developed to solve the heat
Available online 28 January 2012 conduction problem with phase change. In contrast to previous explicit algorithms, the latent heat
source term was treated implicitly in the energy equation, avoiding iteration steps and improving
Keywords: the formulation stability and efficiency. The Bhatnagar–Gross–Krook (BGK) approximation with a
Lattice Boltzmann method D2Q9 lattice was applied and different boundary conditions including Dirichlet and Neumann boundary
Heat transfer conditions were considered. The developed model was tested by solving a one-dimensional melting
Phase change
problem for a pure metal, and one and two-dimensional solidification problems for a binary alloy.
Solidification
The results of the LBM solution were compared with analytical and finite element solutions and a good
consistency was observed. Considering the special capabilities that LBM offers, like local characteristic,
and inherent parallel structure, the developed model is an interesting alternative to traditional contin-
uum models.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction these special capabilities, LBM has attracted the attention of many
researchers and scientists.
The lattice Boltzmann method (LBM) is a relatively new Compu- During the recent years, the application of LBM has been ex-
tational Fluid Dynamics (CFD) technique for solving flow and ther- tended to many areas of fluid and thermal sciences. Heat conduc-
mal problems. While the traditional numerical methods are based tion with phase change is one of the challenging problems that has
on the discretization of conservation equations of continuum many applications in various fields of science and engineering,
mechanics, LBM relies on the solution of a minimal form of Boltz- particularly in metallurgical processes associated with phase
mann kinetic equation for a group of fictive particles in a discret- change like casting, solidification, solid-state phase transforma-
ized domain. The fictive particles stream across the lattice along tions and many other material processes.
the links connecting neighboring lattice sites, and then undergo Wolf-Gladrow [5] was one of the first to develop an LB formula-
collisions upon arrival at a lattice site. For simulating physical phe- tion for diffusion. De Fabritiis et al. [6] developed a thermal model
nomena, the collisions among particles must obey suitable physical for solid–liquid phase change problems by considering different
laws. The fundamental idea of the LBM is to construct kinetic mod- particles for solid and liquid phases. van der Sman et al. [7] devel-
els that incorporate the physics of microscopic processes so that oped a one-dimensional LB model for simulation of heat and mass
the macroscopic averaged properties obey the desired laws. More transport in packed cut flowers. Miller et al. [8] developed a lattice
details about the LB method can be found in the reference books Boltzmann model for anisotropic crystal growth from melt with
published on LBM [1–4]. enhanced computational capabilities. They used similar particles
LBM is an excellent tool for simulation of mass and energy trans- for different phases along with a phase field scheme. Semma
port phenomena. Since the 1990s, LBM has been utilized for solving et al. [9] adopted LBM to solve melting and solidification problems.
a wide variety of transport problems in science and engineering. The They used two distribution functions for fluid flow and heat transfer
major advantages of LBM in comparison with traditional CFD meth- simulations. The phase interface was traced by using partial or prob-
ods consist of simple implementation, capability for simulating abilistic bounce back approach. Jiaung et al. [10] were the first
highly complex geometries and boundaries, computational effi- researchers who introduced an extended lattice Boltzmann equa-
ciency and inherent parallel-processing structure. Considering tion governed by the heat conduction equation in conjunction with
enthalpy method. Chatterjee and Chakraborty [11–13] and Chatter-
jee [14,15] published a series of papers on modeling solid-liquid
⇑ Corresponding author. Address: Mechanical Engineering Department, Mail Stop
phase transition problems using LBM with enthalpy approach. It
9552, 210 Carpenter Building, Mississippi State University, MS 39762, USA. Tel.: +1
662 3251201; fax: +1 662 325 7223. should be noted that explicit approaches have been used in all
E-mail address: felicelli@me.msstate.edu (S.D. Felicelli). studies that employed the enthalpy formulation for phase change

0017-9310/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijheatmasstransfer.2012.01.018
M. Eshraghi, S.D. Felicelli / International Journal of Heat and Mass Transfer 55 (2012) 2420–2428 2421

Nomenclature

Cp specific heat x, y coordinate


c lattice speed X(t) interface location at time t
ei discrete lattice velocity in direction i
fl volume fraction of liquid Greek symbols
gi particle distribution function in direction i a thermal diffusivity
g eq
i equilibrium particle distribution function in direction i x relaxation parameter
h heat transfer coefficient q density
H enthalpy Dfl change in the volume fraction of liquid
k thermal conductivity DH latent enthalpy
L latent heat of fusion Dx lattice spacing in x direction
q_ heat flux Dt time step
St Stefan number U energy source term
t time k solution to the transcendental equation
T temperature Xi(g) rate of local change in the distribution function in direc-
T0 initial temperature tion i due to collisions
TB boundary temperature
Tl liquidus temperature in the binary alloy Subscripts
Tm melting temperature in the pure metal i direction i in a lattice
Ts solidus temperature in the binary alloy l liquid phase
Tsurface surface temperature s solid phase
T1 ambient temperature
wi weight coefficient

calculations [10–15], in which iterations are needed in order to deal 2.2. Lattice Boltzmann model
with the latent heat source term.
In this work, an alternative approach of simulation of heat Lattice Boltzmann models are simpler than the original Boltz-
conduction problem with phase change by using the lattice Boltz- mann equation. The domain is discretized into a number of pseudo
mann method is introduced. While an explicit approach was used particles located on the nodes of the lattice and time is discretized
in previous studies, a novel implicit formulation was adopted for into some distinct steps. There are a few possibilities for spatial
the latent heat source term. The Bhatnagar–Gross–Krook (BGK) position of the particles. One of the most well-known LBM lattices
[16] approximation with a D2Q9 lattice was applied and different is D2Q9 which has two dimensions and nine velocities. Fig. 1
boundary conditions including Dirichlet and Neumann boundary shows the D2Q9 lattice structure which is used in this study. The
conditions were tested. The developed model was utilized to sim- discrete velocities of D2Q9 lattice in 2D Cartesian direction are
ulate the heat conduction during phase change in materials with determined as
constant transition temperature and materials with transition 8
temperature range. The obtained results were compared with the < ð0; 0Þ
> i¼0
results of analytical solutions and other numerical models and ei ¼ ðcos½ði  1Þp=2; sin½ði  1Þp=2Þc i¼14 ð4Þ
>
: pffiffiffiffiffiffi
good consistency was observed. ðcos½ð2i  9Þp=4; sin½ð2i  9Þp=4Þ 2c i ¼ 5  8

where c = Dx/Dt is lattice speed, Dx is lattice spacing and Dt is time


2. Numerical formulation
step.
2.1. Continuum formulation

The Fourier heat conduction equation with phase change can be


written as
@ @
ðqC p TÞ ¼ r:ðkrTÞ þ ðqDHÞ ð1Þ
@t @t
where q, Cp, and k are density, specific heat, and heat conductivity,
respectively. DH is the amount of heat released due to phase
change.
For constant thermo-physical properties, the equation can be
simplified to
@T
¼ ar2 T  U ð2Þ
@t
where a is the heat diffusivity and U is a source term calculated as
@fl L
U¼ ð3Þ
@t C p
where L is the phase change latent heat and fl is the volume fraction Fig. 1. Schematic diagram showing D2Q9 lattice and the unknown distribution
of liquid. functions at west wall and northwest corner.
2422 M. Eshraghi, S.D. Felicelli / International Journal of Heat and Mass Transfer 55 (2012) 2420–2428

The distribution function, gi(x, t), is defined as the probability of Tðx; t þ DtÞ  T s Tðx; tÞ  T s
Dfl ¼ fl ðx; t þ DtÞ  fl ðx; tÞ ¼ 
finding a particle moving in direction i. Then, the macroscopic tem- Tl  Ts Tl  Ts
perature, T(x, t), can be calculated as Tðx; t þ DtÞ  Tðx; tÞ
¼
Tl  Ts
X
8
Tðx; tÞ ¼ g i ðx; tÞ ð5Þ 1 X8
¼ ½g ðx; t þ DtÞ  g i ðx; tÞ ð16Þ
i¼0
T l  T s i¼0 i
The kinetic equation for the distribution functions without an exter-
nal source term can be written as By substituting Eq. (16) into Eq. (14) and then into Eq. (12) and reor-
dering the collision term for a specific particle, the resultant equa-
@ t g i ðx; tÞ þ ei :rg i ðx; tÞ ¼ Xi ðgÞ ð6Þ tion will be

where ei is the velocity in direction i and Xi(g) represents the rate of wi L X8

local change in the distribution function due to collisions. According g i ðx; t þ DtÞ þ g ðx; t þ DtÞ
C p ðT l  T s Þ j¼0 j
to BGK (Bhatnagar–Gross–Krook) approximation, this term can be
expressed as wi L X8
¼ ð1  xDtÞg i ðx; tÞ þ xDtg eq
i ðx; tÞ þ g ðx; tÞ ð17Þ
C p ðT l  T s Þ j¼0 j
Xi ðgÞ ¼ x½g i ðx; tÞ  g eq
i ðx; tÞ ð7Þ
By substituting the temperature from Eq. (5) the collision term can
where x is a relaxation parameter and g eq
i ðx; tÞ is the equilibrium
be defined as
distribution function which can be described as
wi L X8
g eq
i ðx; tÞ ¼ wi Tðx; tÞ ð8Þ g i ðx; t þ DtÞ þ g ðx; t þ DtÞ
C p ðT l  T s Þ j¼0 j
where wi is the weight factor in direction i and for different direc-
wi L
tions of the D2Q9 lattice is given as follows ¼ ð1  xDtÞg i ðx; tÞ þ xDtg eq
i ðx; tÞ þ Tðx; tÞ ð18Þ
C p ðT l  T s Þ
8
< 4=9
> i¼0
By solving for gi(x, t + Dt), the new distribution functions can be
wi ¼ 1=9 i¼14 ð9Þ
>
: implicitly determined from data of the previous time step without
1=36 i ¼ 5  8 iterations.
Note that the formulation stated by Eq. (18) is not equivalent to
After discretization, Eq. (6) can be summarized as the classical modified capacitance method which would have re-
sulted by directly substituting Eq. (15) into Eqs. (2) and (3). An
g i ðx þ ei Dt; t þ DtÞ  g i ðx; tÞ ¼ xDtðg i ðx; tÞ  g eq
i ðx; tÞÞ ð10Þ
LB discretization of Eq. (1) with the source term treated as an addi-
where the LHS can be considered as streaming term and the RHS tional capacitance yields an inefficient and poorly accurate algo-
can be considered as collision term. Collision in LBM terminology rithm, something that was verified with several test simulations.
means relaxation towards the equilibrium. The collision term for This is due to the introduction of dissimilar diffusivities between
a specific particle can be written as regular and interface cells and the fact that the latent heat is not
properly weighted among the distribution functions, as is done in
g i ðx; t þ DtÞ ¼ ð1  xDtÞg i ðx; tÞ þ xDtg eq
i ðx; tÞ ð11Þ Eq. (12). Observe also that although the implicit treatment of the
latent heat term was facilitated by the simple form of Eq. (15), a
By analogy with the continuum formulation, a source term can be similar approach can be followed for more general dependencies
added to the equation of the fraction of liquid including the treatment of solute transport.
In these cases, it is possible to express at least part of the fraction of
g i ðx; t þ DtÞ  g i ðx; tÞ ¼ xDtðg i ðx; tÞ  g eq
i ðx; tÞÞ  Dtwi U ð12Þ liquid relation in terms of the temperature by using the liquidus
line of the phase diagram and the solute conservation equations,
It is proved that Eq. (12) can reproduce the continuum Eq. (2) using as was done for example in Ref. [17] in the context of a mixture
the Chapman–Enskog expansion. For this to be true, the thermal formulation.
diffusivity, a, in the D2Q9 lattice should be set as follows [10]
 
c2 2 2.2.1. LBM boundary conditions
a¼  Dt ð13Þ Defining consistent boundary conditions is crucial in LBM and
6 x
many studies have been done to find the appropriate ways to apply
In the discretized form, U is defined by the following equation various types of boundary conditions in lattice Boltzmann simula-
tions. Here two different types of thermal boundary conditions
L Df l including Dirichlet and Neumann conditions will be discussed.
U¼ ð14Þ
C p Dt
2.2.1.1. Dirichlet boundary conditions. For boundaries aligned paral-
where Dfl is the change in volume fraction of liquid during the time
lel to the coordinate axes, the distribution functions are not known
step Dt. The volume fraction of liquid is calculated with the lever
in three directions. For example g1, g5 and g8 are unknown at the
rule formula:
west wall as it is shown in Fig. 1. The unknown distribution func-
T  Ts tions can be determined using the other known distribution func-
fl ¼ ð15Þ tions and the known constant temperature boundary condition. It
Tl  Ts
is assumed that the unknown distribution functions are of the form
where Tl and Ts are liquidus and solidus temperatures respectively. gi = wig0 , where g0 is the residual temperature needed to satisfy the
Considering Eq. (5), the change in the volume fraction of liquid, Dfl, constant temperature at the boundary [3]. According to Eq. (5), the
can be written as prescribed temperature on the west wall can be written as:
M. Eshraghi, S.D. Felicelli / International Journal of Heat and Mass Transfer 55 (2012) 2420–2428 2423

X
8 step. This is the property which makes the method very appropriate
TB ¼ g i ¼ g 0 þ g 2 þ g 3 þ g 4 þ g 6 þ g 7 þ g 0 ðw1 þ w5 þ w8 Þ ð19Þ for parallelization purposes.
i¼0

Then the residual temperature can be calculated as 3. Verification and discussion


T B  ðg 0 þ g 2 þ g 3 þ g 4 þ g 6 þ g 7 Þ
g0 ¼ In order to investigate the applicability and accuracy of the
w1 þ w5 þ w8
model, three different problems including one-dimensional melt-
¼ 6½T B  ðg 0 þ g 2 þ g 3 þ g 4 þ g 6 þ g 7 Þ ð20Þ
ing of a pure metal, and one-dimensional and two-dimensional
Finally, the unknown distribution functions can be determined solidification of a binary alloy are solved and the obtained results
using the weighting factors are compared with analytical and FEM results.
1 0 1 0 1 0
g 1 ¼ w1 g 0 ¼ g; g 5 ¼ w5 g 0 ¼ g; g 8 ¼ w8 g 0 ¼ g ð21Þ 3.1. One-dimensional melting of a pure metal
9 36 36
Melting of a semi-infinite slab of pure aluminum is simulated in
2.2.1.2. Neumann boundary condition. The Neumann boundary con- one-dimension. The thermo-physical properties of pure aluminum
dition requires prescribing the heat flux or applying a convective are listed in Table 1 and the schematic of the problem is shown in
heat transfer coefficient. In one dimension, this can be written as Fig. 2. The slab is initially assumed to be in the uniform temperature
@T of 600 °C. Then, it is assumed that a constant temperature of 750 °C
k ¼ q_ þ hðT surface  T 1 Þ ð22Þ is imposed to the left wall. The problem is considered one-dimen-
@x
sional since the side walls are insulated and heat transfer occurs
where q_ is the prescribed heat flux, h is the heat transfer coefficient, only in one dimension. In order to compare the numerical results
Tsurface is the surface temperature, T1 is the ambient temperature with the available analytical solutions, the length of the slab is con-
and @T@x
is the temperature gradient. Considering the known temper- sidered 10 cm which is long enough to satisfy the semi-infinity
ature gradient at the surface, the temperature at the boundary can assumption for the considered times and locations. The lattice spac-
be determined using a three-point finite difference scheme [18] ing, Dx, and the relaxation parameter, x, were respectively adopted

2 @T  4 1 as 0.1 mm and 1.0 for all simulations.
Tðx0 Þ ¼ þ Tðx0 þ DxÞ  Tðx0 þ 2DxÞ ð23Þ
3 @x x¼x0 3 3 For melting of a semi-infinite slab, initially at a uniform temper-
ature T0 6 Tm, by imposing a constant temperature TB > Tm on
where x0 corresponds to the location of the boundary. Then, the un- the face x = 0, the analytical solution is given as follows [19]. The
known distribution functions at the boundary can be calculated solid–liquid interface location has the form
using the same scheme used to apply Dirichlet boundary condition. pffiffiffiffiffi
It should be noted that a similar procedure can be employed to XðtÞ ¼ 2k at ð25Þ
impose the boundary conditions at the corners considering the fact Temperature in the liquid region 0 < x < X(t), t > 0 is given by
that the distributions functions are unknown in five directions at
the corners (see Fig. 1). ðT B  T m Þðerfð2pxffiffiffi
at
ffiÞÞ
Tðx; tÞ ¼ T B  ð26Þ
erfk
2.2.2. Numerical procedure and temperature in the solid region x > X(t), t > 0 can be calculated
The sequence in which the numerical calculations are done is as
very important. Typically, the geometry, physical properties, initial
and boundary conditions and LBM parameters, including lattice ðT m  T 0 Þðerfcð2pxffiffiffi
at
ffiÞÞ
Tðx; tÞ ¼ T 0 þ ð27Þ
spacing and relaxation time should be defined at first. The equilib- erfcðkÞ
rium distribution functions are initially obtained using the pre-
scribed temperature according to Eq. (8) and initial distribution Here k is the solution to the transcendental equation
functions, gi(x, 0), are defined as St l St s pffiffiffiffi
2
 2
¼k p ð28Þ
g i ðx; 0Þ ¼ g eq
i ðx; 0Þ ð24Þ expðk ÞerfðkÞ expðk ÞerfcðkÞ

Then, the following steps are repeated in each time step. where
C p ðT B  T m Þ
1. Apply boundary conditions. Stl ¼ ð29Þ
L
2. Compute equilibrium distribution functions using Eq. (8).
3. Collision: and
If T > Tl or T < Ts; calculate relaxation using Eq. (11). C p ðT m  T 0 Þ
If Ts < T < Tl; calculate relaxation using Eq. (18). Sts ¼ ð30Þ
L
4. Calculate temperature using Eq. (5).
5. Streaming. Fig. 3 shows the thermal histories of the points located at 1, 5, 10
and 20 mm from the hot wall. The melting process at the points
close to the hot wall begins shortly after the boundary conditions
are imposed. It can be seen that the heating rate becomes slower
It should be noted that in contrast to previous explicit methods, as the distance from the hot wall increases. A change in the slope
the distribution functions in the collision step can be implicitly of the curves can be seen at the melting temperature. The cause
calculated by solving a small (9  9) system of linear equations at of this change is the latent heat released due to phase transforma-
each node, and iterations and convergence criterion are not needed. tion. However, because of high heating rates at the points close to
Another important point is that except for the streaming step, all the hot wall, the slope change is not considerable at these points.
other steps are completely local and particles do not interact with The temperature profiles at different times are illustrated in
the adjacent particles. During streaming step, the particles only Fig. 4. Again, the slope change due to phase change is observed.
interact with their nearest neighboring particles in the streaming A good consistency can be seen between LBM and analytical results
2424 M. Eshraghi, S.D. Felicelli / International Journal of Heat and Mass Transfer 55 (2012) 2420–2428

Table 1
Thermo-physical properties of the studied materials.

Material Density (kg/m3) Conductivity (W/m °C) Specific heat (J/kg °C) Latent heat (J/kg) Liquidus temperature (°C) Solidus temperature (°C)
3
Pure Al 2698.9 210.0 900.0 386.9  10 660.0 660.0
Al–3wt%Cu 2475.0 30.0 500.0 271.2  103 652.0 596.0

Fig. 2. Schematic of the one-dimensional melting of the pure material showing


coordinates, initial and boundary conditions, liquid and solid zones, and the
interface position.

Fig. 5. LBM and analytical solutions of the interfacial position for the one-
dimensional pure metal melting problem.

Fig. 3. Comparison between LBM and analytical thermal histories of the points Fig. 6. Schematic of the one-dimensional solidification of an Al–3%Cu alloy showing
located at 1, 5, 10, and 20 mm from the hot wall for the one-dimensional pure metal coordinates, initial and boundary conditions, liquid, solid and mushy zones, and the
melting problem. interfacial positions.

Fig. 4. Comparison between LBM and analytical solutions for temperature profiles Fig. 7. Comparison between LBM and FEM cooling curves of the points located at 1,
after 1, 5, 20, and 60 s for the one-dimensional pure metal melting problem. 10, and 25 mm from the cold wall for the one-dimensional binary alloy solidifi-
cation problem.

in both thermal histories and temperature profiles which indicates


the accuracy of the LBM model. It should be noted that since Fig. 5 compares the LBM and analytical solutions for the inter-
the model used for LBM simulations is not actually semi-infinite, facial position. Again, very good agreement is shown between
the results gradually diverge from analytical solution as either time LBM and analytical solutions which suggests that the current LB
or distance from the hot wall increases. However, even in this model can precisely solve the one-dimensional phase change prob-
condition, the maximum relative error at long times and distances lem. The relative difference between FEM and LBM results was less
was less than 2%. than 0.2%.
M. Eshraghi, S.D. Felicelli / International Journal of Heat and Mass Transfer 55 (2012) 2420–2428 2425

3.2. One-dimensional solidification of a binary alloy

Solidification of a binary alloy in one dimension was simulated.


It should be noted that this solution is not practically correct
because solute transport which is a major mechanism in solidifica-
tion of binary alloys has been neglected. However, since phase
change occurs in a temperature range instead of a temperature
point, which is similar to what happens in binary alloys, an
Al–Cu binary alloy was considered for the study. Table 1 shows
the thermo-physical properties of the binary alloy considered in
this study and Fig. 6 illustrates the schematic of the problem.
The slab is initially at the uniform temperature T0 = 700 °C. Then
a constant heat flux q_ = 150,000 W/m2 is imposed on the left wall.
The length of the slab is 50 mm and the end wall is insulated. Here
phase change occurs in a range of temperature instead of a single
Fig. 8. Comparison between LBM and FEM solutions for temperature profiles after temperature. This means that there is a so called mushy zone
1, 5, 20, and 60 s for the one-dimensional binary alloy solidification problem. which is neither completely solid nor completely liquid and con-
tains a mixture of solid and liquid phases.
LBM results for this case are verified against FEM solutions. FEM
simulations are carried out using the commercial software ANSYS
10.0. The quadrilateral bilinear elements PLANE55 were used for
meshing the model. The mesh size was selected fine enough to
yield converged stable results. An enthalpy formulation is used
to simulate phase change phenomenon using FEM.
Solidification starts from the left cold wall and progresses
through the slab. Fig. 7 shows the cooling history of different points
located at 1, 10, and 25 mm from the hot wall. Unlike the previous
problem in which the phase change latent heat was released in a
single temperature, here it is released in a range of temperature.
Hence a sharp change in the slope of the curve cannot be seen here,
but a gradual slope change can still be observed.
Fig. 8 shows the temperature profiles at different times after the
boundary conditions are imposed. As the time increases, the tem-
perature decreases throughout the slab and the solidified front
progresses. Very good agreement can be observed between LBM
Fig. 9. LBM and FEM solutions of the liquid/mushy zone (L/M) and mushy zone/ and FEM results in cooling curves and temperature profiles.
solid (M/S) interfacial positions for the one-dimensional binary alloy solidification LBM and FEM solutions for liquid/mushy zone (L/M) and mushy
problem.
zone/solid (M/S) interfacial positions are compared in Fig. 9. The
L/M curve is sharper than M/S curve which implies that the speed
of L/M interface is faster than M/S interface. Since the end wall is
insulated, both interfaces are coincident after a time around
300 s. The excellent consistency observed between LBM and FEM
results indicates the accuracy of the model for predicting phase
change behavior in one-dimensional solidification problems when
phase change occurs in a range of temperature.

Fig. 10. Schematic of the two-dimensional solidification of the Al–3wt%Cu binary Fig. 11. Comparison between LBM and FEM cooling curve of the point located at the
alloy showing coordinates, dimensions, initial and boundary conditions. center of the model for the two-dimensional binary alloy solidification problem.
2426 M. Eshraghi, S.D. Felicelli / International Journal of Heat and Mass Transfer 55 (2012) 2420–2428

on east, south, and west walls respectively. A convective heat trans-


fer condition with the coefficient h = 100 W/m2 °C and ambient
temperature T1 = 25 °C is also considered for the north wall.
Fig. 10 shows the schematic of this problem. Again, the LBM solution
is verified against an FEM solution obtained with ANSYS software.
Fig. 11 shows the cooling history of a point located at the center
of the slab. The slope change due to phase transformation in the
mushy zone is well distinguishable in this case. As the temperature
reaches the liquidus temperature, the cooling rate decreases owing
to latent heat and once the phase change is completed at solidus
temperature, the cooling rate increases. A very good agreement is
shown between LBM and FEM solutions. The relative difference
in this case was less than 0.15%.
The temperature profiles along x and y axes are depicted in
Fig. 12 at different times after the process begins. Even though
the boundary conditions are non-symmetric, the developed model
has been able to precisely capture the temperature variations in
the slab.
Fig. 13 illustrates the color contours comparing LBM and FEM
solutions for temperature distribution after 15, 45, 90 and 120 s.
Cooling process starts from the cold boundary wall. The minimum
temperature is observed at the southeastern corner where the
highest heat transfer happens. The hottest spot is initially located
at the center of the slab, but it migrates to the northwestern corner
where the mildest heat transfer happens.
Fig. 14 shows the color contours for solidified fractions at differ-
ent times. It can be seen that solidification starts from cold bound-
ary walls and progresses through the center. After 15 s, there is still
a completely liquid region at the center. At t = 90 s, some com-
pletely solidified regions are observed at the corners while all other
regions are mushy. After 90 s, most of the slab is solidified except
Fig. 12. Comparison between LBM and FEM temperature profiles after 15, 45, 90,
for a small region at the hot spot which contains less than 40% li-
120 s for the two-dimensional solidification problem. The profiles are shown along quid. The contours at t = 120 s show that the slab is entirely solid
(a) x-axis, and (b) y-axis. at this time. An excellent consistency is observed between LBM
and FEM contours. The agreement between LBM and FEM contour
3.3. Two-dimensional solidification of a binary alloy lines at the corners is also noteworthy. Considering the complicated
boundary conditions applied in this simulation, this consistency
The material considered in this case is similar to the previous indicates the ability and reliability of the current model for pre-
case but the boundary conditions are more complex and non-sym- cisely solving heat conduction problem with phase change in
metric creating a two dimensional problem. Consider an infinite two-dimensions.
slab with a square cross section of 40  40 mm2. Since the slab is
infinite, heat transfer can be neglected in the longitudinal (z) direc- 3.4. Comparison of computational performance in LBM and FEM
tion and the problem is assumed two-dimensional. The slab is
initially at the uniform temperature T0 = 700 °C. Then, constant heat In order to compare the computational efficiency of LBM and
fluxes equal to 120,000, 100,000 and 80,000 W/m2 are imposed FEM, the problem described in Section 3.3 was solved for different

Fig. 13. Color contours comparing LBM and FEM temperature distributions at different times for the two-dimensional solidification problem.
M. Eshraghi, S.D. Felicelli / International Journal of Heat and Mass Transfer 55 (2012) 2420–2428 2427

Fig. 14. Color contours comparing LBM and FEM solid percentages at different times for the two-dimensional solidification problem.

domain sizes using both LBM and FEM. A lattice spacing of LBM can be parallelized and scaled much better than FEM. This fea-
Dx = 1 mm with relaxation parameter x = 1 was used for all simu- ture is being exploited in a work in progress by the authors.
lations in this part. Identical time step, grid size, initial and bound-
ary conditions were adopted for both LBM and FEM cases and all
4. Conclusions
simulations were run for 60 s on a laptop computer with a 2 GHz
Core 2 Duo CPU and 1 GB of RAM. Bilinear quadrilateral elements
A new algorithm to solve the LB equation for the heat conduc-
and a fast iterative solver were used for the FEM calculations, in or-
tion problem with solid/liquid phase change was developed. While
der to favor its performance. Four different square domains with
previous works used explicit schemes, the current model uses an
20, 50, 100, and 150 grids in each side were examined. Fig. 15 com-
implicit scheme to deal with the latent heat source term of the en-
pares the CPU times of LBM and FEM models with respect to the
ergy equation. The Bhatnagar–Gross–Krook (BGK) approximation
total number of cells/elements in the domain. The results indicate
with a D2Q9 lattice was applied and different boundary conditions
that the present LBM model offers much better computational per-
including Dirichlet and Neumann boundary conditions were con-
formance than FEM. The efficiency of the LBM becomes more obvi-
sidered. Three validation examples including one-dimensional
ous as the domain size increases. The reason is that unlike FEM,
melting of pure Al, and one-dimensional and two-dimensional
LBM is a local method and does not require the assembly of a large
solidification of Al–3wt%Cu alloy were solved. A very good agree-
global matrix and solution of system of equations that grows with
ment between LBM, analytical, and FEM solutions was found for
domain size.
all examples when comparing thermal histories, temperature pro-
It should be noted that the real advantages of LBM appear when
files and interfacial locations. Even for non-symmetric mixed
convection effects are incorporated, something that is demon-
boundary conditions, a very good accuracy was demonstrated. In
strated in [20]. Another important computational advantage of
addition, CPU time comparisons demonstrated that the current
LBM is that, due to locality, the needed communication and pas-
LBM model outperforms FEM in computational efficiency. It should
sage of information between processors reduces significantly and
be noted that while most previous works used fictitious physical
properties, real material properties were employed in this study.
Solving this problem with real material properties using an explicit
approach is cause of convergence issues, requiring a finer mesh
size, smaller time steps, more iterations and higher computational
costs. From this point of view, the implicit scheme developed in
this work is computationally more efficient than previous explicit
schemes. On the other hand, implicit methods require an extra
computation (solving the system of equation), which for this case
is a small system of 9  9 equations. Note also that using D2Q5 lat-
tice is enough for energy and mass transport simulations which
will make the presented model even more efficient, because a
smaller system of 5  5 equations is needed to be solved. However,
whether one should use an explicit or implicit method depends
upon the problem to be solved, as for some problems limitation
in the time step and mesh size is not the main issue. Considering
the special capabilities of LBM, like simplicity of implementation,
stability, accuracy, local characteristic, and inherent parallel struc-
ture, the proposed model offers a great potential for simulating
large scale heat and mass transfer phenomena incorporating phase
Fig. 15. Running time of LBM and FEM models for different domain sizes. transformations.
2428 M. Eshraghi, S.D. Felicelli / International Journal of Heat and Mass Transfer 55 (2012) 2420–2428

Acknowledgements [9] E. Semma, M. El Ganaoui, R. Bennacer, A.A. Mohamad, Investigation of flows in


solidification by lattice Boltzmann method, Int. J. Therm. Sci. 47 (3) (2008)
201–208.
This work was funded by the National Science Foundation [10] W.-S. Jiaung, J.-R. Ho, C.-P. Kuo, Lattice Boltzmann method for the heat
through Grant No. CBET-0931801. The authors also appreciate conduction problem with phase change, Numer. Heat Transfer, Part B 39
(2001) 167–187.
the sponsorship of the Center for Advanced Vehicular Systems
[11] D. Chatterjee, S. Chakraborty, An enthalpy-based lattice Boltzmann model for
(CAVS) in Mississippi State University. Useful discussions with diffusion dominated solid–liquid phase transformation, Phys. Lett. A 341 (1–4)
Dr. Liang Wang at CAVS are gratefully acknowledged. (2005) 320–330.
[12] D. Chatterjee, S. Chakraborty, A hybrid lattice Boltzmann model for solid–
liquid phase transition in presence of fluid flow, Phys. Lett. A 351 (4–5) (2006)
359–367.
[13] D. Chatterjee, S. Chakraborty, An enthalpy-source based lattice Boltzmann
References model for conduction dominated phase change of pure substances, Int. J.
Therm. Sci. 47 (2008) 552–559.
[1] D. Wolf-Gladrow, Lattice-Gas Cellular Automata and Lattice Boltzmann [14] D. Chatterjee, An enthalpy-based thermal lattice Boltzmann model for non-
Models: An Introduction, Springer-Verlag, Berlin–Heidelberg, 2000. isothermal systems, Europhys. Lett. 86 (1) (2009) 14004.
[2] S. Succi, The Lattice Boltzmann Equation for Fluid Dynamics and Beyond, [15] D. Chatterjee, Lattice Boltzmann simulation of incompressible transport
Numerical Mathematics and Scientific Computation, Oxford University Press, phenomena in macroscopic solidification processes, Numer. Heat Transfer,
2001. Part B: Fundam. 58 (1) (2010) 55–72.
[3] Michael C. Sukop, Daniel T. Thorne, Lattice Boltzmann Modeling; An [16] J. Bhatnagar, E.P. Gross, M.K. Krook, A model for collision processes in gases: I.
Introduction for Geoscientists and Engineers, Springer-Verlag, Berlin– Small amplitude processes in charged and neutral one-component system,
Heidelberg, 2006. Phys. Rev. 94 (3) (1954) 511–525.
[4] A. Mohamad, Lattice Boltzmann Method Fundamentals and Engineering [17] S.D. Felicelli, J.C. Heinrich, D.R. Poirier, Finite element analysis of directional
Applications with Computer Codes, Springer, London, 2011. solidification of multicomponent alloys, Int. J. Numer. Methods Fluids 27
[5] D. Wolf-Gladrow, A lattice Boltzmann equation for diffusion, J. Stat. Phys. 79 (1998) 207–227.
(5–6) (1995) 1023–1032. [18] C.-H. Liu, K.-H. Lin, H.-C. Mai, C.-A. Lin, Thermal boundary conditions for
[6] G. De Fabritiis, A. Mancini, D. Mansutti, S. Succi, Mesoscopic models of liquid/ thermal lattice Boltzmann simulation, Comput. Math. Appl. 59 (7) (2010)
solid phase transitions, Int. J. Mod. Phys. C 9 (8) (1998) 1405–1415. 2178–2193.
[7] R.G.M. van der Sman, M.H. Ernst, A.C. Berkenbosch, Lattice Boltzmann scheme [19] V. Alexiades, A.D. Solomon, Mathematical Modeling of Melting and Freezing
for cooling of packed cut flowers, Int. J. Heat Mass Transfer 43 (4) (2000) 577– Processes, Taylor & Francis, London, 1993.
587. [20] H. Yin, S.D. Felicelli, L. Wang, Simulation of dendritic microstructure with
[8] W. Miller, S. Succi, A lattice Boltzmann model for anisotropic crystal growth lattice Boltzmann and cellular automaton methods, Acta Mater. 59 (2011)
from melt, J. Stat. Phys. 112 (2002) 173–186. 3124–3136.

También podría gustarte