lattice boltzmann model for melting with natural convection · lattice boltzmann model for melting...

12
Lattice Boltzmann model for melting with natural convection Christian Huber a, * , Andrea Parmigiani b , Bastien Chopard b , Michael Manga c , Olivier Bachmann d a Department of Earth and Planetary Science, University of California – Berkeley, 307 McCone Hall 4767, Berkeley, CA 94720-4767, USA b Computer Science Department, University of Geneva, 24, Rue du Général Dufour, 1211 Geneva 4, Switzerland c Department of Earth and Planetary Science, University of California – Berkeley, 177 McCone Hall 4767, Berkeley, CA 94720-4767, USA d Department of Earth and Space Science, University of Washington, Johnson Hall 070, Seattle WA 98195-1310, USA article info Article history: Received 17 October 2007 Received in revised form 1 May 2008 Accepted 5 May 2008 Available online xxxx Keywords: Lattice Boltzmann Heat transfer Melting Convection abstract We develop a lattice Boltzmann method to couple thermal convection and pure-substance melting. The transition from conduction-dominated heat transfer to fully-developed convection is analyzed and scal- ing laws and previous numerical results are reproduced by our numerical method. We also investigate the limit in which thermal inertia (high Stefan number) cannot be neglected. We use our results to extend the scaling relations obtained at low Stefan number and establish the correlation between the melting front propagation and the Stefan number for fully-developed convection. We conclude by showing that the model presented here is particularly well-suited to study convection melting in geometrically com- plex media with many applications in geosciences. Ó 2008 Elsevier Inc. All rights reserved. 1. Introduction Melting caused by natural convection occurs in many settings, from large-scale phenomena in geosciences to small-scale indus- trial processes during alloy solidification and crystal growth. Mod- els of convection melting need to account for thermally-driven flow coupled with a moving interface where latent heat is either absorbed (melting) or released (solidification). The interplay be- tween the fluid flow and the moving boundary leads to a complex dynamical behavior, as the position of the solid liquid interface be- comes one of the unknowns of the problem (Jany and Bejan, 1988). The variables of interest are the melting front position and the Nusselt number, which describe, respectively, the evolution of the geometry of the system and the heat transfer. The moving boundary problem usually requires complex numerical schemes such as front tracking methods (Bertrand et al., 1999), adaptative grid methods (Mencinger, 2004), level set methods (Tan and Zab- aras, 2006), phase field approaches (Boettinger et al., 2002) or vol- ume-of-fluid methods (Hirt and Nichols, 1981). Natural convection melting has been investigated by numerous experimental (Bénard et al., 2006; Wolff and Viskanta, 1987; Dong et al., 1991; Hirata et al., 1993; Wang et al., 1999), theoretical (Viskanta, 1982; Viskanta, 1985; Jany and Bejan, 1988; Zhang and Bejan, 1989) and numerical studies (Webb and Viskanta, 1986; Bertrand et al., 1999; Mencinger, 2004; Usmani et al., 1992; Chatterjee and Chakraborty, 2005; Javierre et al., 2006). The development of appropriate scaling laws (e.g. Jany and Bejan, 1988) and powerful computational methods have significantly im- proved the understanding of the convection melting processes. Heat transfer correlations for the Nusselt number have been devel- oped, but the range of values for the key dimensionless groups (Rayleigh, Prandtl and Stefan numbers) over which the correlations have been tested remains limited. The lattice Boltzmann method, developed over the last two dec- ades, provides a powerful alternative approach for studying con- vection involving phase changes. Compared with classic computational fluid dynamics methods, it offers two significant advantages. First, no-slip boundary conditions in complex geome- tries can be implemented through simple local rules (Chopard and Droz, 1998). Second, the main part of the algorithm is purely local, making the LB method straightforward to parallelize even though not necessarily more efficient than for example the ‘‘continuum” fi- nite difference method (Nourgaliev et al., 2000, 2003). In this study we develop a Lattice Boltzmann (LB) method to model pure substance conduction and convection melting. We start by introducing the mathematical description of the problem together with previously established scaling laws (Jany and Bejan, 1988) for the convection case. We then introduce the LB method, with extensions for the thermal model (using a multiple distribu- tion approach) and the phase transition (using a modified version of Jiaung et al. (2001) algorithm). In Section 4, we compare the re- sults of our model with analytical solutions (for the conduction case) and with the scaling laws obtained by Jany and Bejan 0142-727X/$ - see front matter Ó 2008 Elsevier Inc. All rights reserved. doi:10.1016/j.ijheatfluidflow.2008.05.002 * Corresponding author. E-mail addresses: [email protected] (C. Huber), andrea.parmigiani@ terre.unige.ch (A. Parmigiani), [email protected] (B. Chopard), man- [email protected] (M. Manga), [email protected] (O. Bachmann). International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx Contents lists available at ScienceDirect International Journal of Heat and Fluid Flow journal homepage: www.elsevier.com/locate/ijhff ARTICLE IN PRESS Please cite this article in press as: Huber, C. et al., Lattice Boltzmann model for melting with natural convection, Int. J. Heat Fluid Flow (2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

Upload: others

Post on 29-Sep-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx

ARTICLE IN PRESS

Contents lists available at ScienceDirect

International Journal of Heat and Fluid Flow

journal homepage: www.elsevier .com/locate / i jhf f

Lattice Boltzmann model for melting with natural convection

Christian Huber a,*, Andrea Parmigiani b, Bastien Chopard b, Michael Manga c, Olivier Bachmann d

a Department of Earth and Planetary Science, University of California – Berkeley, 307 McCone Hall 4767, Berkeley, CA 94720-4767, USAb Computer Science Department, University of Geneva, 24, Rue du Général Dufour, 1211 Geneva 4, Switzerlandc Department of Earth and Planetary Science, University of California – Berkeley, 177 McCone Hall 4767, Berkeley, CA 94720-4767, USAd Department of Earth and Space Science, University of Washington, Johnson Hall 070, Seattle WA 98195-1310, USA

a r t i c l e i n f o

Article history:Received 17 October 2007Received in revised form 1 May 2008Accepted 5 May 2008Available online xxxx

Keywords:Lattice BoltzmannHeat transferMeltingConvection

0142-727X/$ - see front matter � 2008 Elsevier Inc. Adoi:10.1016/j.ijheatfluidflow.2008.05.002

* Corresponding author.E-mail addresses: [email protected] (C.

terre.unige.ch (A. Parmigiani), [email protected]@seismo.berkeley.edu (M. Manga), [email protected]

Please cite this article in press as: Huber,(2008), doi:10.1016/j.ijheatfluidflow.2008

a b s t r a c t

We develop a lattice Boltzmann method to couple thermal convection and pure-substance melting. Thetransition from conduction-dominated heat transfer to fully-developed convection is analyzed and scal-ing laws and previous numerical results are reproduced by our numerical method. We also investigatethe limit in which thermal inertia (high Stefan number) cannot be neglected. We use our results to extendthe scaling relations obtained at low Stefan number and establish the correlation between the meltingfront propagation and the Stefan number for fully-developed convection. We conclude by showing thatthe model presented here is particularly well-suited to study convection melting in geometrically com-plex media with many applications in geosciences.

� 2008 Elsevier Inc. All rights reserved.

1. Introduction

Melting caused by natural convection occurs in many settings,from large-scale phenomena in geosciences to small-scale indus-trial processes during alloy solidification and crystal growth. Mod-els of convection melting need to account for thermally-drivenflow coupled with a moving interface where latent heat is eitherabsorbed (melting) or released (solidification). The interplay be-tween the fluid flow and the moving boundary leads to a complexdynamical behavior, as the position of the solid liquid interface be-comes one of the unknowns of the problem (Jany and Bejan, 1988).The variables of interest are the melting front position and theNusselt number, which describe, respectively, the evolution ofthe geometry of the system and the heat transfer. The movingboundary problem usually requires complex numerical schemessuch as front tracking methods (Bertrand et al., 1999), adaptativegrid methods (Mencinger, 2004), level set methods (Tan and Zab-aras, 2006), phase field approaches (Boettinger et al., 2002) or vol-ume-of-fluid methods (Hirt and Nichols, 1981).

Natural convection melting has been investigated by numerousexperimental (Bénard et al., 2006; Wolff and Viskanta, 1987; Donget al., 1991; Hirata et al., 1993; Wang et al., 1999), theoretical(Viskanta, 1982; Viskanta, 1985; Jany and Bejan, 1988; Zhangand Bejan, 1989) and numerical studies (Webb and Viskanta,

ll rights reserved.

Huber), [email protected] (B. Chopard), man-shington.edu (O. Bachmann).

C. et al., Lattice Boltzmann.05.002

1986; Bertrand et al., 1999; Mencinger, 2004; Usmani et al.,1992; Chatterjee and Chakraborty, 2005; Javierre et al., 2006).The development of appropriate scaling laws (e.g. Jany and Bejan,1988) and powerful computational methods have significantly im-proved the understanding of the convection melting processes.Heat transfer correlations for the Nusselt number have been devel-oped, but the range of values for the key dimensionless groups(Rayleigh, Prandtl and Stefan numbers) over which the correlationshave been tested remains limited.

The lattice Boltzmann method, developed over the last two dec-ades, provides a powerful alternative approach for studying con-vection involving phase changes. Compared with classiccomputational fluid dynamics methods, it offers two significantadvantages. First, no-slip boundary conditions in complex geome-tries can be implemented through simple local rules (Chopard andDroz, 1998). Second, the main part of the algorithm is purely local,making the LB method straightforward to parallelize even thoughnot necessarily more efficient than for example the ‘‘continuum” fi-nite difference method (Nourgaliev et al., 2000, 2003).

In this study we develop a Lattice Boltzmann (LB) method tomodel pure substance conduction and convection melting. Westart by introducing the mathematical description of the problemtogether with previously established scaling laws (Jany and Bejan,1988) for the convection case. We then introduce the LB method,with extensions for the thermal model (using a multiple distribu-tion approach) and the phase transition (using a modified versionof Jiaung et al. (2001) algorithm). In Section 4, we compare the re-sults of our model with analytical solutions (for the conductioncase) and with the scaling laws obtained by Jany and Bejan

model for melting with natural convection, Int. J. Heat Fluid Flow

Page 2: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

solidH liquid

z

x

m

T=melting temperature

T=T >melting temperature1

l

x (t)

z

i

h

s

H

l

Solid

liquid

ii iii ivi

1

adiabatic and no–slip

T=T

adiabatic and no–slip

T=Tmno–slip

adiabatic and no–slip

Fig. 1. (a) Schematic representation of the conduction melting problem. (b)Schematic representation of the different stages of the convection melting problem.The solid is initially at the melting temperature Tm: (i) boundary conditions; (ii)conduction and convection; (iii) convection dominated; (iv) after the melting frontreached the right boundary.

2 C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx

ARTICLE IN PRESS

(1988) for the convection problem. We extend their scaling laws tohigh Stefan numbers (i.e. where thermal inertia is non-negligible)for the Nusselt number and the position of the melting front. Final-ly, we show that the model developed here is able to handle non-idealized problems with complex geometries such as porous mediawith no additional complications.

2. Review of pure substance melting

The problem of half space conduction melting with homoge-neous, isotropic thermal diffusivity, has been solved analyticallyin 1860 by Neumann. Heat transfer in the liquid is given by

oTot¼ jr2T; ð1Þ

where T is temperature, and j the thermal diffusivity. Nomencla-ture is summarized in Table 1. At the melt-solid boundary, whenthe solid is kept at the melting temperature, the energy balance re-quires that

joTox

� �x¼xm

¼ Lf

cdxm

dt: ð2Þ

Here, c is the heat capacity, xm is the position of the meltingfront and Lf the latent heat of fusion (see Fig. 1a). The problem isoften recast by separating enthalpy in a sensible heat and a latentheat term (Faghri and Zhang, 2006),

oTot¼ jr2T � Lf

cofl

ot; ð3Þ

where fl is the melt fraction; or in dimensionless form,

oT�

ot�¼ r2T� � 1

Stofl

ot�; ð4Þ

where the dimensionless scalings are T� ¼ ðT � T0Þ=ðT1 � T0Þ;x� ¼ x=l, l is the natural length scale of the system; t� ¼ tj=l2. T1,T0 are, respectively, the temperature of the superheat applied with-in the liquid and the initial temperature of the solid (or, in the caseof no undercooling, the melting temperature). St is the Stefan num-ber, St ¼ cðT1 � T0Þ=Lf . It is important to note that Eq. (1) togetherwith Eq. (2) are exactly equivalent with Eq. (3). Whereas the firstversion ensures thermal energy balance at the interface through aboundary condition, the second introduces the latent heat contribu-

Table 1Nomenclature

c specific heat of the liquid phasec1 fitting parameter for Nu—Fo correlationc2 fitting parameter for Nu—Fo correlationEn enthalpyEns enthalpy of the solid phase at the melting temperatureEnl enthalpy of the liquid phase at the melting temperaturefl liquid fractionFo Fourier number ¼ jt=H2

g acceleration due to gravityH height of enclosurel width of the enclosureLf latent heat of fusionNu Nusselt numberp pressurePr Prandtl number ¼ m=jRa Rayleigh number ¼ gbðT1 � T0ÞH3=ðjmÞSt Stefan number ¼ cðT1 � T0Þ=Lf

T1 temperature of the superheat applied to the liquidT0 temperature of the solid (here melting temperature)b coefficient of thermal expansionj thermal diffusivity of the fluidk coefficient depending on St (Eq. (5))m kinematic viscosity

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

tion in the diffusion equation as a sink term. As a result, when thelatter is discretized, a careful choice of time step is required in orderto ensure energy conservation at the interface otherwise the sinkterm can prevail over the heat transport and lead to non-physicalbehavior (see Section 3.2).

For a system starting with zero undercooling (the solid is ini-tially at the melting temperature), Neumann found that the solu-tion in the liquid half-space is given by

Tðx; tÞ ¼ T1 � ðT1 � T0Þerfðx=ð2

ffiffiffiffiffiffijtpÞÞ

erfðkÞ ; for 0 6 x 6 xmðtÞ;

xmðtÞ ¼ 2kffiffiffiffiffiffijtp

;

k expðk2ÞerfðkÞ ¼ Stffiffiffiffipp ;

ð5Þ

where xm is the melting front position.Using the same set of dimensional scales, the system of equa-

tions for convection melting, in the case of Newtonian incompress-ible fluids under the Boussinesq approximation (and neglectingviscous heating), can be written as Bodenschatz et al. (2000)

r � u� ¼ 0; ð6Þou�

ot�þ u� � ðru�Þ ¼ �rp� þ Prr2u� � PrRabT�; ð7Þ

oT�

ot�þ u� � rT� ¼ r2T� � 1

Stofl

ot�; ð8Þ

where Pr ¼ m=j is the Prandtl number and Ra ¼ gbDTl3=ðjmÞ is the

Rayleigh number, m, b and g are, respectively, the kinematic viscos-ity, the thermal expansion and the acceleration due to gravity and lis an appropriate length scale for the problem (l will correspond tothe height of the enclosure for the convection melting problem).

Owing to their non-linear behavior the equations for convectionmelting cannot be solved analytically and require numerical inte-gration. However, for simple geometries and flow behavior (lowReynolds numbers, Boussinesq approximation), Jany and Bejan(1988) showed that using appropriate representative lengthscales,it is possible to infer scaling relationships for the evolution of thesystem. These scaling laws have been successfully tested bothnumerically (Webb and Viskanta, 1986; Jany and Bejan, 1988; Ber-trand et al., 1999) and experimentally (Bénard et al., 2006; Wolff

model for melting with natural convection, Int. J. Heat Fluid Flow

Page 3: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx 3

ARTICLE IN PRESS

and Viskanta, 1987; Gobin and Bénard, 1992; Wang et al., 1999).However, owing to the difficulty of experiments in satisfying per-fectly isothermal or adiabatic walls and no density change inducedby the phase transition, we decide to validate our numerical modelwith the scaling laws of Jany and Bejan (1988) and the benchmarkexercise of Bertrand et al. (1999). Fig. 1b shows the boundary con-ditions and the different stages for the melting of a 2D rectangularcavity of height H.

The first stage is dominated by heat conduction melting,although owing to the vertical geometry of the cavity, sluggishconvection will take place because of the horizontal temperaturedifferences in the thin fluid layer. During this stage, Jany and Bejanshowed that the Nusselt number is

Nu / h�1=2 þ Ra h3=2; ð9Þ

where h ¼ FoSt, Fo ¼ jt=l2 is the Fourier number, andRa ¼ gbðT1 � T0ÞH3=ðjmÞ. This stage ends when the convection zonethickness z (see Fig. 1b-iii) extends to the height of the cavity. Thedimensionless time h1 at which this occurs was found to be relatedto the Rayleigh number according to the law

h1 / Ra�1=2: ð10Þ

Finally, the scaling predicts that the Nusselt number will evolve to aminimum

Numin / Ra1=4; ð11Þ

occurring close to the end of this regime.In the convection regime (z ¼ H), the Nusselt number reaches a

plateau at

Nu / Ra1=4: ð12Þ

The height-averaged position of the melting front sav and thedimensionless time at which melting reaches the right boundaryof the domain h2 are given by

sav ¼1H

Z H

0xm dz / HRa1=4h ð13Þ

and

h2 ¼lH

Ra�1=4; ð14Þ

where l is the horizontal dimension of the cavity. Finally, Jany andBejan (1988) showed that the scaling results listed above are validat high Prandtl number (Pr > Oð1Þ) and proposed to rescaleRa! RaPr when Pr < 1 in every scaling law.

3. Lattice Boltzmann (LB) model

In the LB model, the fluid is described by quantities fi represent-ing the particle density distributions in the ith velocity direction ofthe lattice;

fi ¼ fiðx; tÞ; i ¼ 0; . . . ;M; ð15Þ

where x is the position on the lattice and t the time. M representsthe number of velocity directions of the particles at each node ofthe lattice. The volumetric mass q and the momentum qu are de-fined as particle velocity moments of the distribution function (Fris-ch et al., 1986)

q ¼XM

i¼0

fi; qu ¼XM

i¼1

fiei; ð16Þ

where ei is the local particle velocity. In the single relaxation time(or LBGK) model, the equation for fi is (Qian et al., 1992; Chopardand Droz, 1998)

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

fiðxþ ei; t þ 1Þ ¼ fiðx; tÞ �1sðfiðx; tÞ � f eq

i ðx; tÞÞ: ð17Þ

In this relation, the relaxation parameter s expresses the rate atwhich the local particle distribution relaxes to the local equilibriumstate f eq

i .Let us consider a two-dimensional regular grid with a nine-

velocity lattice (the so-called D2Q9 topology) (Wolf-Gladrow,2000). The velocities ei are given by

ei ¼

ð0;0Þ i ¼ 0;ðcosðði� 1Þp=2Þ; sinðði� 1Þp=2ÞÞ i ¼ 1;2;3;4;ffiffiffi

2pðcosðð2ði� 5Þ þ 1Þp=4Þ; sinðð2ði� 5Þ þ 1Þp=4ÞÞ

i ¼ 5;6;7;8:

8>>><>>>:

ð18Þ

In this case, the equilibrium distribution function may be written as

f eqi ¼ qwi 1þ 3ðei � uÞ þ

92ðei � uÞ2 �

32

u � u� �

ð19Þ

with w0 ¼ 16=36, wi ¼ 4=36 for i ¼ 1;2;3;4 and wi ¼ 1=36 fori ¼ 5;6;7;8. With the appropriate equilibrium distribution, anddefining the speed of sound as c2

s ¼ c2=3, we can recover the conti-nuity and Navier–Stokes equations through a Chapman–Enskogexpansion (Frisch et al., 1986; Qian et al., 1992; Chopard and Droz,1998). Accordingly, the pressure p is identified with p ¼ qc2

s and thekinematic viscosity (in lattice units) is defined by

m ¼ c2s dtðs� 0:5Þ ð20Þ

where dt is the timestep. The positivity of the viscosity requiress > 0:5. For more details about the algorithm, the reader is referredto lattice Boltzmann textbooks (Chopard and Droz, 1998; Succi,2001; Wolf-Gladrow, 2000).

3.1. Thermal lattice Boltzmann

When the latent heat term is neglected, Eq. (8) becomes thestandard advection–diffusion equation for temperature

oT�

ot�þ u� � rT� ¼ r2T�: ð21Þ

The existing LB models developed for thermal flows can be classi-fied into four categories: multispeed (MS), entropic, hybrid andmulti-distribution function (MDF) models.

The MS approach (McNamara and Alder, 1993a,b) is a straight-forward extension of the isothermal LBGK model in which only oneparticle distribution function fi is used to recover the entire set ofthermo-hydrodynamics equations. Density, momentum and inter-nal energy conservation are obtained from the moments of the dis-tribution fi. For computational efficiency, the lattice topology ischosen so as to minimize the number of discrete velocities forwhich the LB model recovers the macroscopic dynamics of interest.In order to recover the desired set of equations (conservation laws),a thermal MS model needs a larger number of lattice velocitiescompared to an isothermal LBGK model. One of the obvious restric-tions of using a single LBGK model for thermal problems is theimpossibility of varying the Prandtl number.

Prasianakis and Karlin (2007) have developed a single distribu-tion thermal LB scheme with variable Prandtl number on standardlattices (e.g., D2Q9 for a 2D model). They use an entropic latticeBoltzmann scheme (Ansumali and Karlin, 2002) for which a correc-tion term recovers the Navier–Stokes equation (in the compress-ible regime), the advection–diffusion equations for temperature,and also provides a free parameter for tuning the Prandtl number.However, the correction term is non-local. Hybrid models couplethe flow field calculated with a conventional LB scheme with finitedifferences or finite elements to solve for the heat equation (Lalle-mand and Luo, 2003; Mezrhab et al., 2004).

model for melting with natural convection, Int. J. Heat Fluid Flow

Page 4: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

4 C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx

ARTICLE IN PRESS

In the present study, we choose the MDF approach (Shan, 1997;Guo et al., 2002) to model natural convection. The thermal advec-tion–diffusion Eq. (21) is solved by introducing a second distribu-tion function (gi) whose evolution is also described by a LBGKdynamic. The two distributions are coupled through the buoyancyterm for fi, and the equilibrium distribution for gi,

fiðxþ ei; t þ 1Þ ¼ fiðx; tÞ �1sðfiðx; tÞ � f eq

i ðx; tÞÞ þ g

� eiPr RaTðtÞ=DT; ð22Þ

giðxþ vi; t þ 1Þ ¼ giðx; tÞ �1shðgiðx; tÞ � geq

i ðx; tÞÞ: ð23Þ

The LBGK model for gi is similar to Eq. (17) but a simplified equilib-rium distribution function can be used:

geqi ¼ TwT

i ½1þ1

C2sh

ðvi � uÞ�; ð24Þ

where vi and wTi are the associated lattice velocities and weights,

and u is the macroscopic fluid flow velocity. Eqs. (23) and (24)are first order approximations, therefore, correcting terms have tobe used to make it second order accurate (Lätt, 2007). The MDF ap-proach allows us to use two different lattices for the two distribu-tion functions. For the evolution of gi, given its simplifiedequilibrium distribution function, a D2Q5 lattice is preferred. Inthe D2Q5 topology, the velocities vi are

vi ¼ð0;0Þ i ¼ 0;ðcosðði� 1Þp=2Þ; sinðði� 1Þp=2ÞÞ i ¼ 1;2;3;4:

�ð25Þ

The associated weights wTi are wT

0 ¼ 1=3;wTi ¼ 1=6 for i ¼ 1;2;3;4.

At each lattice node, the macroscopic temperature is defined as

T ¼X4

i¼0

gi ð26Þ

and the thermal diffusivity (in lattice units) is related to the relax-ation time

j ¼ C2sh dtðsh � 0:5Þ: ð27Þ

The MDF approach offers the possibility to vary the Prandtl numberby adjusting the two relaxation times s and sh

Pr ¼ ð2s� 1Þð2sh � 1Þ : ð28Þ

3.2. Phase change with lattice Boltzmann

Different LB approaches have been proposed for solid–liquidphase transitions. They can be grouped in two methods: (1) thephase-field method using the theory of Ginzburg–Landau (Millerand Succi, 2002; Rasin et al., 2005; Medvedev and Kassner,2005); (2) enthalpy-based methods (Jiaung et al., 2001; Chatterjeeand Chakraborty, 2005). We use a sightly modified version of theJiaung et al. (2001) melting scheme for the conduction case, usinga D2Q5 topology. Jiaung et al. use an iterative enthalpy-basedmethod to solve for both the temperature and melt fraction fieldsat each time step. The melting term is introduced as a source (crys-tallization) or sink (melting) term in the collision step. In summary,at the timestep n, iteration k, the macroscopic temperature iscalculated

Tn;k ¼X4

i¼0

gn;ki ; ð29Þ

where Tn;k � Tkðt ¼ nÞ. The local enthalpy is obtained by

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

Enn;k ¼ cTn;k þ Lf fn;k�1l ð30Þ

with the liquid fraction fl of the previous iteration. Finally, the en-thalpy is used to linearly interpolate the melt fraction

f n;kl ¼

0 Enn;k < Ens ¼ cTm;

Enn;k�EnsEnl�Ens

Ens 6 Enn;k6 Ens þ Lf ;

1 Enn;k > Ens þ Lf :

8>><>>:

ð31Þ

The collision is then calculated

gn;kþ1i ðxþeiÞ¼ giðxÞ�

1sh

giðxÞ�geqi ðxÞ

� �wi

Lf

cf n;kl ðxÞ� f n�1

l ðxÞ �

;

ð32Þ

until the temperature and the melt fraction field converge to withina set tolerance. The timestep nþ 1 is calculated using the same pro-cedure. Jiaung et al. (2001) obtain accurate solutions forOð10�1Þ 6 St 6 Oð1Þ and a relaxation time sh ¼ 1. For the sake ofefficiency, we modified the scheme and set the number of iterationsto k ¼ 18n at the expense of the accuracy. However, we performedtests revealing that, over the range of parameters used in this study,setting k ¼ 1 has negligible effects, but becomes valuable for thecomputationally intensive convection melting problems.

As the Jiaung et al. (2001) scheme is a LB version of Eq. (3), weshow that special attention must be given to the choice of param-eters used in the numerical calculations. Using the total time of therun tmax as the timescale, we get a new non-dimensional version ofthe differential equation

oT�

ot�¼ jtmax

l2 r2T� � 1St

ofl

ot�: ð33Þ

Eq. (33) shows that for a melting event the temperature evolu-tion is subject to a source term (the heat transported from the hotwall) and a sink term (the heat consumed by the phase change).When jtmax=l2

< cnstSt�1, the sink term dominates and the tem-perature can locally decrease below the initial solid temperature,in contradiction with the phase change associated with pure sub-stance melting. Therefore the choice of spatial and temporal dis-cretization is important (will modify jtmax=l2), especially at lowthermal diffusivity. Accordingly, we modified the collision step toensure no heat transfer in parts of the domain where the enthalpyhas remained equal to the initial enthalpy of the solid. This is doneby applying the collision as in Jiaung et al. (2001) for every site xwhere the enthalpy EnðxÞ > cT ini and using bounceback for theremaining sites. This is equivalent to forcing the boundary condi-tion of Eq. (2) in the Jiaung et al. scheme. As a result, the temper-ature field obtained numerically is in better agreement withtheory, especially when using small relaxation times (0:5 <sh < 1) and a small number of time steps.

On the other hand, if the Stefan number is high compared withjtmax=l2, there will be a limit at which the scheme cannot solve theevolution of the melting front. The condition is that the meltingfront velocity dxm=dt does not exceed the lattice speed dx=dt.Using Eqs. (5) and (27), in lattice units, this condition translates to

sh < 0:5þ 31k2 ; ð34Þ

which gives approximately sh < 3:5 and sh < 48:5 for, respectively,St ¼ 10 and St ¼ 0:1. In Eq. (34), k is given by Eq. (5).

For the convection melting problem, we develop a new schemebased on the thermal lattice Boltzmann model described in Section3.1. The two distributions are coupled through the buoyancy forceand the equilibrium distribution geq

i (for the advection term). Anadditional coupling arises from the melting scheme through themelting front position determined by the temperature scheme.

model for melting with natural convection, Int. J. Heat Fluid Flow

Page 5: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Pos

ition

of t

he m

eltin

g fr

ont

Normalized time

κ =0.033κ =0.366

κ =0.7

-0.2

0

0.2

0.4

0.6

0.8

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Nor

mal

ized

tem

pera

ture

T*

Normalized distance x/L

κ =0.033κ =0.366

κ =0.7

Fig. 2. Comparison of the evolution of the melting front position (a), and thetemperature distribution after 10,000 iterations (b) in the conduction case for threedifferent thermal diffusivities and their respective analytical solutions (lines) fromEq. (5), with Stefan number St ¼ 1.

C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx 5

ARTICLE IN PRESS

When a site becomes liquid, the distribution functions fi’s are ini-tialized with the equilibrium values with a reference density anda zero velocity (because of the no-slip boundary condition). Wealso used a velocity extrapolated from the neighboring fluid sites,and found no notable differences. The collision step for the fluiddistribution is modified accordingly, and we perform the standardcollision of Eq. (17) at every site where the liquid fraction is above0.5 (this choice is arbitrary and represents the limit at which mostof the voxel of the given site is in the fluid state) and bouncebackeverywhere else. For the wall boundary conditions, bouncebackis applied on the four walls for the fluid distribution. For the sakeof simplicity, especially for complex geometries, we use on-gridbounceback and define the solid–liquid interface to be located atmid-grid to recover second order accuracy. Upper, lower and rightwalls are adiabatic (no heat flow out) and are implementedthrough the bounceback of the temperature distribution (Kumaret al., 1999; van der Sman, 2004, 2006). We tested the bouncebackof the temperature distribution against a second order finite differ-ence approximation for the adiabatic boundaries and we found agood agreement (no difference for the Nusselt number and withina percent of each other for the average melting front position whenit reached the half width of the enclosure). Finally, the left wall isset at a fixed temperature T1 by setting the only unknown g1 to

g1 ¼ T1 � ðg0 þ g2 þ g3 þ g4Þ ð35Þ

for all sites on the left wall.We emphasize that the choice jsolid ¼ jliquid is not a necessary

condition and that the viscosity can be set to be a function of thetemperature by rescaling the relaxation time of the fluid distribu-tion s accordingly (Guo and Zhao, 2005). The two latter extensionswill not be discussed in this study, but are important for actualapplications. Lastly, the scheme presented here is not fully incom-pressible; for convection melting problems at higher Mach num-bers, the incompressible model of Guo et al. (2000) is required.

4. Results and discussion

In this section we present the results of both conduction andconvection melting with our lattice Boltzmann model and comparethem with analytical solutions (for the conduction case) and withthe scaling obtained by Jany and Bejan (1988).

4.1. Conduction melting

The different results obtained for conduction melting (the Ste-fan problem) are summarized in Figs. 2–4. We use a 100 � 5 gridfor all runs. The geometry and boundary conditions are illustratedin Fig. 1a. Fig. 2a shows the melting front position as function oftime for three different thermal diffusivities j ¼ 0:033;0:36 and0.7 in lattice units, corresponding to relaxation times s of 0.6, 1.6and 2.6, respectively, with a fixed Stefan number of St ¼ 1. Fig.2b shows the temperature distributions at the end of each run(corresponding to 10,000 time steps) shown in Fig. 2a. The resultof each of these simulations is compared with their respective ana-lytical solutions (Eq. (5)). For both melting front position and tem-perature, the numerical and analytical results are in goodagreement. However, as expected, the fit is better at low thermaldiffusivity for our explicit scheme, which, for a fixed grid spacing,is equivalent to smaller time steps, or, for fixed time steps, is equiv-alent to a better spatial resolution.

Fig. 3 shows the effect of the temporal resolution and the Stefannumber on the results obtained for both melting front position andtemperature. For each figure, we chose three relaxation times (0.6,1.6 and 2,6) and ran for the same dimensionless time t� ¼ tj=l2.Again, the results are in good agreement with the analytical solu-tions. However, the runs corresponding to the highest thermal dif-

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

fusivity (j ¼ 0:7), i.e. the runs with the worst temporal resolution(at least 10 times fewer iterations), resolve the interface less accu-rately. The variation of the Stefan number over two orders of mag-nitude does not change the aptitude of the explicit algorithm tosolve the Stefan problem.

We can define the error for the temperature distribution andthe melting front position to be, respectively

dT ¼Pnx�1

i¼0 jTðxi; tÞ � Tthðx ¼ xi; tÞjPnx�1i¼0 T thðx ¼ xi; tÞ

and ð36Þ

dxmðtÞ ¼jxmðtÞ � xth

mðtÞjxth

m ðtÞ; ð37Þ

where the superscript ‘th’ refers to the analytical solution.Fig. 4a shows the evolution of the relative melting front position

error with time for a Stefan number of 0.13. The error in the melt-ing front position decreases rapidly to reach of few percents aftert� ¼ 0:2. As expected, the results get better with better temporalresolution. The same feature is observed for the error in the tem-perature distribution dT of Fig. 4b. Finally, we investigate thedependence of the error in the Stefan number. Figs. 4c and d showthat the error on the melting front position decreases with increas-ing St. This result is expected as the possible error introduced bythe explicit algorithm to solve for the non-linear term of Eq. (4)is multiplied by St�1 and consequently decreases as the Stefan

model for melting with natural convection, Int. J. Heat Fluid Flow

Page 6: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0 0.4 0.8 1.2 1.6 2

x m/l

t*=tκ/l 2

analyticκ =0.033κ =0.366κ =0.7

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0 0.1 0.2 0.3 0.4

x m/l

t*=tκ/l 2

analyticκ =0.033κ =0.366κ =0.7

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0 0.02 0.04 0.06 0.08 0.1

x m/l

t*=tκ/l2

analyticκ =0.033κ =0.366κ =0.7

-0.2

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

T*

Normalized distance x/L

analyticκ =0.033κ =0.366κ =0.7

0

0.05

0.7 0.75

T*

-0.2

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1T*

Normalized distance

analyticκ =0.033 κ =0.366κ =0.7

0

0.05

0.75 0.8

T*

-0.2

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

T*

Normalized distance

analyticκ =0.033κ =0.366κ =0.7

0

0.05

0.75 0.8T*

Fig. 3. Comparison of the melting front position and temperature profiles for the same three thermal diffusivities with Stefan number St ¼ 0:13 (first row), St ¼ 1 (secondrow) and St ¼ 10 (third row). Time is made dimensionless in order to compare the sensitivity of the model to the size of the timesteps.

6 C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx

ARTICLE IN PRESS

number increases. Mathematically, in the limit St !1 we recovera simple single phase heat conduction equation from Eq. (4).

4.2. Convection melting

Convection simulations were performed for a square geometry,with the number of lattice points fixed by the value of the desiredRayleigh number and the constraints for stability mentioned inSection 3. The grid sizes range from 75 � 75, for Ra ¼ 5� 104, to250 � 250, for Ra ¼ 6:8� 106. Fig. 5 shows a grid resolution testfor Ra ¼ 25000, Pr ¼ 0:02 and St ¼ 0:01 using three different gridsizes. The three results are in very good agreement with Bertrandet al. (1999).

We compared our results with the benchmark of Bertrand et al.(1999) for the case: Pr ¼ 0:02, St ¼ 0:01, Ra ¼ 2:5� 105. The evolu-tion of the average melting front position and the Nusselt numberare in good agreement with the main trend of results showed inBertrand et al. (1999). The relatively high Nusselt number andoscillatory nature of the results are similar to the results of LeQuéré and Couturier–Sadat listed in Bertrand et al. (1999), wherethey attribute this behavior to the full transient procedure andthe evolution of the circulation cells as melting proceeds.

In this section, we first test our numerical method by recoveringthe scaling relationships and the correlation coefficients for Numin,Nu2, hmin and h2 developed by Jany and Bejan (1988) at both low

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

and high Stefan number 0 6 St 6 10. Then a comparable run tothe one described in Jany and Bejan (1988) (St ¼ 0:1, Pr > Oð1Þ)is used to reproduce the correlations that were obtained. Second,we show the effect of thermal inertia at high St and compare itto the theory of Jany and Bejan (1988). A review of low Prandtlnumber effects can be found in Gobin and Bénard (1992), whereit was shown that the scaling laws, in this case, have a slightdependence on the Prandtl number.

Fig. 6 shows the evolution of the temperature field in the squarecavity as function of time, from the early conduction-dominatedstage (upper left) to the fully developed convection (upper right)and finally the entirely melted state (lower right). This run wasset with Ra ¼ 6:8� 106, Pr ¼ 1 and St ¼ 10. The evolution of themelting front is in qualitative good agreement with the expectedschematic geometries described in Fig. 1b.

For a more quantitative perspective, we can compare the runscomputed for four different Ra ranging from 5� 104 to 6:8� 106

with the theoretical predictions obtained from the scaling relation-ships of Section 2. Fig. 7a shows the evolution of the average meltfront position for the four different Ra numbers and the conductioncase. As expected the evolution at small h is controlled by conduc-tion, until more efficient convection increases the melting rate toreach a linear trend during the fully-developed convection regime.Finally the melting rate slows as the melt front reaches the rightboundary of the domain. The Nusselt number as function of dimen-

model for melting with natural convection, Int. J. Heat Fluid Flow

Page 7: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

0.001

0.01

0.1

1

0.001 0.01 0.1 1

δxm

t* = κt/l2

0.01x-0.48

κ =0.033κ =0.366

κ =0.7

0.001

0.01

0.1

1

0.001 0.01 0.1 1

δT

t* = tκ/l2

0.0025x-0.84

κ =0.033κ =0.366

κ =0.7

0.01

0.1

1

1e-04 0.001 0.01 0.1

δxm

t* = κt/l2

St=0.13St=1

St=10

0.001

0.01

0.1

1

10

1e-05 1e-04 0.001 0.01 0.1δT

t* = κt/l2

St=0.13St=1

St=10

Fig. 4. (a) Relative error on the position of the melting front and on the temperature distribution (b) for the data of Fig. 3 (St ¼ 0:13); (c) Relative error on the position of themelting front and on the temperature distribution (d) for j ¼ 0:366 and St ¼ 0:13, 1 and 10, respectively.

0

5

10

15

20

25

0 0.02 0.04 0.06 0.08 0.1

Nu

50 x 5075 x 75

150 x 150

0

5

10

15

20

25

0 0.02 0.04 0.06 0.08 0.1

Nu

= Fo Stθ= Fo Stθ

Ra=2.5 x 104 ; St=0.01; Pr=0.02 Ra=2.5 x 105 ; St=0.01; Pr=0.02

Fig. 5. (a) Resolution tests using three grid sizes 50 � 50, 75 � 75 and 150 � 150. Ra ¼ 2:5� 104, Pr ¼ 0:02 and St ¼ 0:01. (b) Validation run for Ra ¼ 2:5� 105, Pr ¼ 0:02 andSt ¼ 0:01. The results are similar to the results of Le Quéré and Couturier–Sadat listed in Bertrand et al. (1999). The high Nusselt number and oscillatory behavior of (b) can beexplained by the fully transient procedure and small time steps (Bertrand et al., 1999).

C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx 7

ARTICLE IN PRESS

sionless time is shown in Fig. 7b for the same four runs. The Nusseltnumber is computed as in Jany and Bejan (1988)

Nu ¼Z H

0

oT�

ox�ðx ¼ 0; zÞdz: ð38Þ

The overall shape of the curves display the expected featuresdescribed by Jany and Bejan (1988): Nu / h�1=2 at the early stage,and then a minimum followed by a plateau. It is worth noting thatthe curves here show more time variability than those of Jany andBejan (1988), especially at high Ra, which can be explained by thelow Pr number (=1) used here (higher Reynolds number). We canextract Nu and time of its minimum value and the point at whichthe melting front reaches the right boundary and apply the scalingrelationships described in Eqs. (10), (11), (12) and (14). Fig. 8shows that the numerical results are in very good agreement withthe scaling laws and enable us to extract the constants of propor-tionality. We added the results obtained by Jany and Bejan(1988) for St ¼ 0:1 in Fig. 8b for comparison. The correlations weobtained (within a few percent) at Pr ¼ 1 and St ¼ 10 are listedin Table 2 together with the results of Jany and Bejan (1988).

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

Interestingly, the correlations for the time h reported here aregreater than those of Jany and Bejan (1988) by a factor five, whileNu correlations are in good agreement. From the scaling relations,a possible explanation is that the scaling laws derived for low St haveto be corrected for the relatively large St we used. One way to under-stand this difference is by taking a closer look at the Nu� h scaling atshort times (when conduction is dominating the heat transfer). TheNeumann solution for purely conductive melting leads to

Nu ¼ 1erfðkÞ ðpFoÞ�1=2 ð39Þ

in the limit of small St this can be simplified to

Nu ¼ ð2hÞ�1=2: ð40Þ

A simple calculation shows that Eq. (40) underestimates the temporalevolution of the Nusselt number by a factor of approximately threefor St ¼ 10, and thus the high St will tend to shift Nu to higher h. Forexample, Jany and Bejan (1988) propose the following correlationfor the Nusselt number evolution (at low St):

model for melting with natural convection, Int. J. Heat Fluid Flow

Page 8: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

Fig. 6. Plot of the temperature field for different dimensionless times (h). Ra ¼ 6:8� 106, Pr ¼ 1 and St ¼ 10.

8 C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx

ARTICLE IN PRESS

0

10

20

30

40

50

0 0.2 0.4 0.6 0.8 1

Nu

θ = Fo Stθ = Fo St

Ra=5e4Ra=1.7e5Ra=8.4e5Ra=6.8e6

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1 1.2

s av/l

conductionRa=5e4

Ra=1.7e5Ra=8.4e5Ra=6.8e6

Fig. 7. (a) Average melting front position as function of dimensionless time h for four different numerical runs, Pr ¼ 1 and St ¼ 10 for all. (b) Nusselt number evolution for fourdifferent numerical runs, Pr ¼ 1 and St ¼ 10 for all.

0.1

1

10

100

1000 10000 100000 1e+06 1e+07Ra

Numin/Ra1/4

θmin/Ra-1/2

θ2/Ra-1/4

Nu2/Ra1/4

0.1

1

10

100

1000

10000

0.1 1 10 100St

θmin/Ra-1/2

θ2/Ra-1/4

Numin/Ra1/4

Nu2/Ra1/4

Jany and Bejan(1988)

Fig. 8. Scalings from the numerical modeling. (a) Fixed Pr = 1 and St = 10. (b) Fixed Pr = 1 and Ra ¼ 105. The two fits give hmin=Ra�1=2 ¼ 3:57St1:05 þ 12:3 andh2=Ra�1=4 ¼ 1:1St0:8 þ 2:12.

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann model for melting with natural convection, Int. J. Heat Fluid Flow(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

Page 9: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

0

5

10

15

20

25

30

35

40

0.01 0.1 1

Nu

θ = Fo St

Ra=5e4, Pr=1, St=10Eq. (41) low St correlation

Eq. (42) corrected correlation

Fig. 9. Comparison between numerical results and the correlation function from Eq.(41) and its corrected version for high St. In both cases, the fit was improved byreplacing c2 ¼ 0:0175 by c2 ¼ 0:00175.

Table 2Comparison of the correlations obtained in this study (St ¼ 10) with those of Jany andBejan (St ¼ 0:1)

Numin hmin Nu2 h2

Jany and Bejan (1988) 0.28 Ra1=4 9 Ra�1=2 0.35 Ra1=4 1.8 Ra�1=4

This study St ¼ 10 0.31 Ra1=4 56 Ra�1=2 0.29 Ra1=4 8.7 Ra�1=4

C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx 9

ARTICLE IN PRESS

Nu ¼ ð2hÞ�1=2 þ c1Ra1=4 � ð2hÞ�1=2h i

1þ ðc2Ra3=4h3=2Þnh i1=n

; ð41Þ

where the constant c1 ¼ 0:29 is derived from the correlation of Nu2

with Ra1=4, c2 and n are fitting parameters (they found c2 ¼ 0:0175

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

sav

/l

θ = Fo St

Ra=5e4Ra=1.7e5Ra=8.4e5Ra=6.8e6

0.1

1

10000 100000 1e+06 1e+07

C1

Ra

C1=0.33

Fig. 10. Correlation for the thermal inertia effect using the scaling relation dsav=dh ¼ 00.33 ± 0.02. (b) Ra ¼ 1� 105 and Pr ¼ 1; C1 ranges from 0.33 to 0.46. The slopes of the straconvection stage. (c) C1 as function of Ra for Pr ¼ 1 and St ¼ 10. (d) C1 as function of St

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

and n ¼ �2). Here we propose a better correlation for the evolutionof the Nusselt number by replacing each occurrence of ð2hÞ�1=2 by(39) in Eq. (41):

Nu ¼ 1erfðkÞ ðpFoÞ�1=2

þ c1Ra1=4 � 1erfðkÞ ðpFoÞ�1=2

� �½1þ ðc2Ra3=4h3=2Þn�1=n

: ð42Þ

We find that the fit is much improved (see Fig. 9), suggesting thatthe scalings for hmin and h2 are not exactly independent of St (espe-cially at high St). Other consequences of thermal inertia will be dis-cussed later. To summarize the results at high St number, weobserved that the scaling relations of Section 2 are in good agree-ment with our numerical results, and, the correlations for Numin

and Nu2 are very close to those observed by Jany and Bejan(1988), although the temporal position of these characteristicpoints is shifted to higher values of h which seems to be consistentwith the higher St used in our runs.

In order to validate the results we obtain from our lattice Boltz-mann model, we compare the results of a numerical calculationmade by Jany and Bejan (1988) (estimated from graphs in their pa-per) with Ra ¼ 1� 105, Pr ¼ 50 and St ¼ 0:1 with a run with ourscheme at similar Ra and St and slightly lower Pr (=5). Accordingto the scaling laws (Jany and Bejan, 1988) and experimental results(Gobin and Bénard, 1992), at Pr > 1, the relationship between theNusselt number and the Rayleigh number are independent of thePrandtl number, therefore the difference in Pr number betweenthe two runs should not affect the comparison. Fig. 8b shows theconstants we obtain for the scaling laws at St ¼ 0:1; they are ingood agreement with the results obtained by Jany and Bejan(1988).

Thermal inertia effects due to large superheat (high St number)have been introduced theoretically by Jany and Bejan (1988),

θ = Fo St

0

0.5

1

1.5

2

0 0.1 0.2 0.3 0.4 0.5 0.6

St=0.1St=1St=2St=5

St=10

0.1

1

1 10

St

C1(St)=St-0.06-0.54

:29Ra1=4=ð1þ C1StÞ. (a) St ¼ 10 and Pr ¼ 1; C1 obtained from these correlations isight lines, given by Eq. (44), match the numerical results during the fully-developedfor Ra ¼ 105 and Pr ¼ 1.

model for melting with natural convection, Int. J. Heat Fluid Flow

Page 10: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

10 C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx

ARTICLE IN PRESS

where they derive an expression for the propagation of the averagemelting front position during the fully developed convection stage

dsav

dh¼ c1Ra1=4

1þ C1St; ð43Þ

where c1 ¼ 0:29 and C1 is the constant incorporating thermal iner-tia effects into the melting front propagation. Fig. 10a shows thebest fit from Eq. (43) and the corresponding values for C1 are listedin Fig. 10c as a function of Ra (St ¼ 10). Figs. 10b and d show thedependence of C1 on the Stefan number. The best fits obtained fromEq. (43) for the melting front propagation during the fully convec-tive regime (see Fig. 10b) lead to the values shown in Fig. 10d. Asconjectured by Jany and Bejan (1988), we can confirm that C1 isindependent of the Stefan number over the range of Stefan numberexplored. Therefore, the evolution of the melting front in the fully-developed convection regime is best represented by the equation

Fig. 11. Porous media convection melting. For each h, two plots are shown, the temperepresents the solid). (For interpretation of the references to color in this figure legend,

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

dsav

dh¼ c1Ra1=4

1þ 0:33St; ð44Þ

for 5� 1046 Ra 6 7� 106 and 0:1 6 St 6 10.

Suga (2006) showed that the conventional D2Q5 advection–dif-fusion scheme is stable for the condition C2

x þ C2y 6 2=5 for

1 < sh 6 4, where Cx ¼ uxdt=dx, Cy ¼ uydt=dy are the Courant num-bers in the x and y directions, respectively. For low relaxation times(sh � 0:502), the advection–diffusion scheme is stable formaxðjCxj; jCyjÞ 6 2=5. The additional term accounting for the latentheat released, in our temperature scheme, brings an additionalconstraint on the Stefan number for the stability of the scheme(see Eq. (34)). However, these two stability constraints are notindependent, and stability problems can still occur for large simu-lations at both high Ra and Pr. Stability can be recovered in thesecases by reducing the duration of the time steps at the expenseof computational time. Finally, Suga (2006) showed that the accu-

rature field in the upper row and the melt fraction in the lower row (where bluethe reader is referred to the web version of this article.)

model for melting with natural convection, Int. J. Heat Fluid Flow

Page 11: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx 11

ARTICLE IN PRESS

racy of the D2Q5 model for advection–diffusion problems dependsmostly on the Peclet number (Pe ¼ udx=j), with a good accuracywhen Pe < 10. All the numerical results presented in this studywere well within the stability range of the D2Q5 model and per-formed at Pe < 1. We have tested different grid sizes and foundthat the appropriate grid size is limited by the stability and accu-racy of the scheme at very low relaxation times (� 0:501) for smallgrids and by the computational time for very large grids. Withinthis range, which depends on Ra, Pr and St, we find a good agree-ment between the different runs.

The model we describe in this study can solve for melting in por-ous media without modification. Convection melting in complexgeometries (porous media) have many application in geosciences.For example, in volcanic settings, magma chambers host compli-cated dynamics with coexisting solid and liquid phases (as well asgas phases in general). Magmatic systems experience a complextemporal evolution with successive stages of reheating (related toinjection of new magma for example) and cooling, resulting in a cor-responding complex evolution of the local melt fraction by eithermelting or solidification. High energy reheating events are invokedto explain several eruptions of systems with relatively high crystal-linity (around 40–50%) (Bachmann and Bergantz, 2003, 2006).

We apply our lattice Boltzmann convection melting model to asimulation of pure substance melting in a synthetic porous media(porosity of 45%). We use a 400 � 400 grid, the thermal diffusivityand the viscosity are set, respectively, to 1/3 and 1/6 in latticeunits. Results are shown in Fig. 11. The domain is heated from be-low at a fixed temperature (above the melting temperature), theside walls are treated as periodic boundaries (for both temperatureand fluid distributions) and the top wall is set as a no-slip, fixedtemperature (melting temperature) boundary. The simulation isset with Pr ¼ 0:5 and St ¼ 1:3 (arbitrary choice). The systems meltsfrom bottom up until it reaches a critical average height that en-ables large scale convection (at around h ¼ 3:6� 10�2). After thispoint, the melting rate is dominated by convection. The averagepore-size is small enough to block convection beyond the meltingfront. In contrast with the Stefan problem, the melting frontboundary is not flat because of horizontal temperature gradientsdue to the inhomogeneous distribution of the solid.

5. Conclusions

Lattice Boltzmann models offer a simple and powerful alterna-tive to classic numerical methods when dealing with reactive flowin complex geometries. In this study, we present an explicit latticeBoltzmann scheme that can solve conduction and convection melt-ing problems over a wide range of dimensionless parameters(especially Ra and St). Results show good agreement with scalinglaws obtained by Jany and Bejan (1988). This approach also allowsus to extend the scaling for the evolution of Nusselt number. Inparticular, our results suggest that the correlation developed byJany and Bejan (1988) for the small St limit can be modified andgeneralized to St > 1. The correlation between the propagation ofthe melting front and the Stefan number during the fully devel-oped convection regime is also established. We also show thatthe model presented here permits, without any subsequent modi-fication, a study of convection melting in porous media.

Acknowledgments

We thank Josef Dufek, Jonas Lätt and Jim Watkins for their help-ful comments and stimulating discussions. The authors thank thereviewers and the editor for their helpful suggestions. This workwas supported by NSF EAR 0439766.

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

References

Ansumali, S., Karlin, I., 2002. Single relaxation time model for entropic latticeBoltzmann methods. Phys. Rev. E 65 (5), 05631.

Bachmann, O., Bergantz, G.W., 2003. Rejuvenation of the Fish Canyon magma body:a window into the evolution of large-volume silicic magma systems. Geology31, 789–792.

Bachmann, O., Bergantz, G.W., 2006. Gas percolation in upper-crustal magmabodies as a mechanism for upward heat advection and rejuvenation of siliciccrystal mushes. J. Volcanol. Geotherm. Res. 149, 85–102.

Bénard, C., Gobin, D., Martinez, F., 2006. Melting in rectangularenclosures:experiments and numerical simulations. J. Heat Transf. 107, 794–803.

Bertrand, O., Binet, B., Combeau, H., Couturier, S., Delannoy, Y., Gobin, D., Lacroix, M.,Le Quéré, P., Médale, M., Mencinger, J., Sadat, H., Vieira, G., 1999. Melting drivenby natural convection. A comparison exercise: first results. Int. J. Therm. Sci. 38,5–26.

Bodenschatz, E., Pesch, W., Ahlers, G., 2000. Recent developments in Rayleigh–Bénard convection. Annu. Rev. Fluid Mech. 32, 709–778.

Boettinger, W.J., Warren, J.A., Beckermann, C., Karma, A., 2002. Phase-fieldsimulation of solidification. Annu. Rev. Mater. Res. 32, 163–194.

Chatterjee, D., Chakraborty, S., 2005. An enthalpy-based lattice Boltzmann modelfor diffusion dominated solid–liquid phase transformation. Phys. Lett. A 341,320–330.

Chopard, B., Droz, M., 1998. Cellular Automata and Modeling of Physical Systems.Monographs and Texts in Statistical Physics. Cambridge University Press. p. 353.

Dong, Z.F., Chen, Z.Q., Wang, Q.J., Ebadian, M.A., 1991. Experimental and analyticalstudy of contact melting in a rectangular cavity. J. Thermophys. 5, 347–354.

Faghri, A., Zhang, Y., 2006. Transport Phenomena in Multiphase Systems. Elsevier. p.1030.

Frisch, U., Hasslacher, B., Pomeau, Y., 1986. Lattice gas automata for the Navier–Stokes equations. Phys. Rev. Lett. 56, 1505–1508.

Gobin, D., Bénard, C., 1992. Melting of metals driven by natural convection in themelt: influence of Prandtl and Rayleigh numbers. J. Heat Transf. 114, 521–524.

Guo, Z., Shi, B., Wang, N., 2000. Lattice BGK model for incompressible Navier–Stokesequation. J. Comput. Phys. 165, 288–306.

Guo, Z., Shi, B., Zheng, C., 2002. A coupled lattice BGK model for the Boussinesqequations. Int. J. Numer. Methods Fluids 39 (5), 325–342.

Guo, Z., Zhao, T.S., 2005. Lattice Boltzmann simulation of natural convection withtemperature-dependent viscosity in a porous cavity. Progr. Comput. Fluid Dyn.5, 110–117.

Hirata, T., Makino, Y., Kaneko, Y., 1993. Analysis of natural convection meltinginside isothermally heated horizontal rectangular capsule. WärmeStoffübertrag. 28, 1–9.

Hirt, C.W., Nichols, B.D., 1981. Volume of fluid (VOF) method for the dynamics offree boundaries. J. Comput. Phys. 39, 201–225.

Javierre, E., Vuik, C., Vermolen, F.J., van der Zwaag, S., 2006. A comparison ofnumerical models for one-dimensional Stefan problems. J. Comput. Appl. Math.192, 445–459.

Jany, P., Bejan, A., 1988. Scaling theory of melting with natural convection in anenclosure. Int. J. Heat Mass Transf. 31, 1221–1235.

Jiaung, W.-S., Ho, J.-R., Kuo, C.-P., 2001. Lattice-Boltzmann method for the heatconduction problem with phase change. Numer. Heat Transf.: Part B 39, 167–187.

Kumar, R., Nivarthi, S.S., Ted Davis, H., Kroll, D.M., Maier, R.S., 1999. Application ofthe lattice-Boltzmann method to study flow and dispersion in channels withand without expansion and contraction geometry.Int. J. Numer. Methods Fluids31, 801–819.

Lallemand, P., Luo, L.-S., 2003. Theory of the lattice Boltzmann method: acoustic andthermal properties in two and three dimensions. Phys. Rev. E 68, 036706.

Lätt, J., 2007. Hydrodynamic Limit of Lattice Boltzmann Equations, PhD Thesis,University of Geneva, p. 102.

McNamara, G., Alder, B., 1993a. Analysis of the lattice Boltzmann treatment ofhydrodynamics. Physica A 194 (5), 218.

McNamara, G., Alder, B., 1993b. Lattice Boltzmann thermohydrodynamics. Phys.Rev. E 47 (4), 2249–2252.

Medvedev, D., Kassner, K., 2005. Lattice Boltzmann scheme for crystal growth inexternal flows. Phys. Rev. E 72, 056703.

Mencinger, J., 2004. Numerical simulation of melting in two-dimensional cavityusing adaptative grid. J. Comput. Phys. 198, 243–264.

Mezrhab, A., Bouzidi, M., Lallemand, P., 2004. Hybrid lattice-Boltzmann finite-difference simulation of convective flows. Comput. Fluids 33, 623–641.

Miller, W., Succi, S., 2002. A lattice Boltzmann model for anisotropic crystal growthfrom melt. J. Stat. Phys. 112, 173–186.

Nourgaliev, R.R., Dinh, T.A., Dalal, D.C., Dinh, T.N., Theofanous, T.G., 2000. MuSiC:multiscale simulation code. modeling of complex fluid–fluid and fluid–solidinteractions using the level-set and ‘ghost fluid’ methodology. Single- andMultiphase Compressible Flows, UCSB-CRSS Research Report, November 20, p.209.

Nourgaliev, R.R., Dinh, T.N., Theofanous, T.G., Joseph, D., 2003. The lattice Boltzmannequation method: theoretical interpretation, numerics and implications. Int. J.Multiphase Flow 29, 117–169.

Prasianakis, N., Karlin, I., 2007. Lattice Boltzmann method for thermal flowsimulation on standard lattices. Phys. Rev. E 76, 016702.

model for melting with natural convection, Int. J. Heat Fluid Flow

Page 12: Lattice Boltzmann model for melting with natural convection · Lattice Boltzmann model for melting with natural convection Christian Hubera,*, Andrea Parmigianib, Bastien Chopardb,

12 C. Huber et al. / International Journal of Heat and Fluid Flow xxx (2008) xxx–xxx

ARTICLE IN PRESS

Qian, Y.H., D’Humières, D., Lallemand, P., 1992. Lattice BGK models for the Navier–Stokes equation. Europhys. Lett. 17, 479–484.

Rasin, I., Miller, W., Succi, S., 2005. Phase-field lattice kinetic scheme for thenumerical simulation of dendritic growth. Phys. Rev. E 72, 066705.

Shan, X., 1997. Simulation of Rayleigh–Benard convection using a lattice Boltzmannmethod. Phys. Rev. E 55 (5), 2780–2788.

Succi, S., 2001. The lattice Boltzmann equation for fluid dynamics and beyond.Numerical Mathematics and Scientific Computation. Oxford University Press. p.288.

Suga, S., 2006. Numerical schemes obtained from lattice Boltzmann equations foradvection diffusion equations. Int. J. Mod. Phys. C 17 (11), 1563–1577.

Tan, L., Zabaras, N., 2006. A level set simulation of dendritic solidification withcombined features of front-tracking and fixed-domain methods. J. Comput.Phys. 211, 36–63.

Usmani, A.S., Lewis, R.W., Seetharamu, K.N., 1992. Finite element modelling ofnatural-convection-controlled change of phase. Int. J. Numer. Methods Fluids14, 1019–1036.

van der Sman, R.G.M., 2004. Diffusion on unstructured triangular grids using latticeBoltzmann. Future Gener. Comput. Syst. 20, 965–971.

Please cite this article in press as: Huber, C. et al., Lattice Boltzmann(2008), doi:10.1016/j.ijheatfluidflow.2008.05.002

van der Sman, R.G.M., 2006. Galilean invariant lattice Boltzmann scheme for naturalconvection on square and rectangular lattices. Phys. Rev. E 74, 026705.

Viskanta, R., 1982. Phase-change heat transfer. In: Lane, G.A. (Ed.), Solar HeatStorage: Latent Heat Materials. CRC Press, Boca Raton, FL, pp. 153–222.

Viskanta, R., 1985. Natural convection melting and solidification. In: Kakac, S., Aung,W., Viskanta, R. (Eds.), Natural Convection: Fundamentals and Applications.Hemisphere, Washington, DC, pp. 845–877.

Wang, Y., Amiri, A., Vafai, K., 1999. An experimental investigation of the meltingprocess in a rectangular enclosure. Int. J. Heat Mass Transf. 42, 3659–3672.

Webb, B.W., Viskanta, R., 1986. Analysis of heat transfer during melting of a puremetal from an isothermal vertical wall. Numer. Heat Transf. 9, 539–558.

Wolff, F., Viskanta, R., 1987. Melting of pure metal from a vertical wall. Exp. HeatTransf. 1, 17–30.

Wolf-Gladrow, D.A., 2000. Lattice-Gas Cellular Automata and Lattice BoltzmannModels: An Introduction. Lecture Notes in Mathematics. Springer. p. 308.

Zhang, Z., Bejan, A., 1989. The problem of time-dependent natural convectionmelting with conduction in the solid. Int. J. Heat Mass Transf. 32, 2447–2457.

model for melting with natural convection, Int. J. Heat Fluid Flow