12
International Journal of Modern Physics B Vol. 22, No. 22 (2008) 3865–3876 c World Scientific Publishing Company SIMPLIFIED FINITE DIFFERENCE THERMAL LATTICE BOLTZMANN METHOD C. S. NOR AZWADI Faculty of Mechanical Engineering, Universiti Teknologi Malaysia, 81310 UTM Skudai, Johor, Malaysia [email protected] T. TANAHASHI Department of Mechanical Engineering, Keio University, Yokohama, Kanagawa, Japan [email protected] Received 5 November 2007 In this paper, a well-known finite difference technique is combined with thermal lattice Boltzmann method to solve 2-dimensional incompressible thermal fluid flow problems. A small number of microvelocity components are applied for the calculation of temperature field. The combination of finite difference with lattice Boltzmann method is found to be an efficient and stable approach for the simulation at high Rayleigh number of natural convection in a square cavity. Keywords : Thermal lattice Boltzmann; finite difference; FDLBM; natural convection flow. 1. Introduction The lattice Boltzmann method (LBM) is considered as an alternative approach to the well-known finite difference, finite element, and finite volume techniques for solving the Navier-Stokes equations. LBM evolved from Lattice Gas Automata, 1 simulates fluid flows by tracking the evolution of the single-particle distribution. Although as a newcomer in numerical scheme, the lattice Boltzmann approach has found recent successes in a host of fluid dynamical problems, including flows in porous media, magnetohydrodynamics, immiscible fluids, and turbulence. How- ever, the simulation of flows with heat transfer turned out to be much more difficult. In general, the current thermal lattice Boltzmann models fall into three cate- gories: the multi-speed approach, 2 the passive scalar approach, 3 and the thermal energy distribution model proposed by He et al. 4 The multi-speed approach uses 3865

SIMPLIFIED FINITE DIFFERENCE THERMAL LATTICE ...azwadi/pdf_publication/ISI3.pdfSeptember 3, 2008 10:0 WSPC/140-IJMPB 04861 International Journal of Modern Physics B Vol. 22, No. 22

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    International Journal of Modern Physics BVol. 22, No. 22 (2008) 3865–3876c© World Scientific Publishing Company

    SIMPLIFIED FINITE DIFFERENCE THERMAL LATTICE

    BOLTZMANN METHOD

    C. S. NOR AZWADI

    Faculty of Mechanical Engineering, Universiti Teknologi Malaysia,

    81310 UTM Skudai, Johor, Malaysia

    [email protected]

    T. TANAHASHI

    Department of Mechanical Engineering, Keio University,

    Yokohama, Kanagawa, Japan

    [email protected]

    Received 5 November 2007

    In this paper, a well-known finite difference technique is combined with thermal latticeBoltzmann method to solve 2-dimensional incompressible thermal fluid flow problems. Asmall number of microvelocity components are applied for the calculation of temperaturefield. The combination of finite difference with lattice Boltzmann method is found to bean efficient and stable approach for the simulation at high Rayleigh number of naturalconvection in a square cavity.

    Keywords: Thermal lattice Boltzmann; finite difference; FDLBM; natural convectionflow.

    1. Introduction

    The lattice Boltzmann method (LBM) is considered as an alternative approach to

    the well-known finite difference, finite element, and finite volume techniques for

    solving the Navier-Stokes equations. LBM evolved from Lattice Gas Automata,1

    simulates fluid flows by tracking the evolution of the single-particle distribution.

    Although as a newcomer in numerical scheme, the lattice Boltzmann approach has

    found recent successes in a host of fluid dynamical problems, including flows in

    porous media, magnetohydrodynamics, immiscible fluids, and turbulence. How-

    ever, the simulation of flows with heat transfer turned out to be much more

    difficult.

    In general, the current thermal lattice Boltzmann models fall into three cate-

    gories: the multi-speed approach,2 the passive scalar approach,3 and the thermal

    energy distribution model proposed by He et al.4 The multi-speed approach uses

    3865

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    3866 C. S. Nor Azwadi & T. Tanahashi

    the same distribution function in defining the macroscopic velocity, pressure, and

    temperature. In addition to mass and momentum, in order to preserve the kinetic

    energy in the collision on each lattice point, this model requires more variations

    of speed than those of the isothermal model and equilibrium distribution function

    usually include higher order velocity terms. However, this model is reported to

    suffer severe numerical instability, and is not computationally efficient.5

    In the passive scalar model, the flow fields (velocity and density) and the tem-

    perature are represented by two different distribution functions. The macroscopic

    temperature is assumed to satisfy the same evolution equation as a passive scale,

    which is advected by the flow velocity, but does not affect the flow field. It has

    been shown that the passive scalar model has better numerical stability than the

    multi-speed model.5

    He et al.4 in their model introduce the internal energy density distribution func-

    tion, which can be derived from the Boltzmann equation. This model is shown to

    be a suitable model for simulating real thermal problems. However, the complicated

    gradient operator term appears in the evolution equation and thus the simplicity

    property of the lattice Boltzmann scheme has been lost.6

    The conventional isothermal and thermal lattice Boltzmann models, however,

    only give second-order accuracy in space and time. Since Luo and He7,8 and Abe9

    demonstrated that the lattice Boltzmann equation is a discretized form of the con-

    tinuous Boltzmann equation and the discretization of physical space is not coupled

    with the discretization of momentum space, any standard numerical technique can

    serve the purpose of solving the discrete Boltzmann equation. It is not surprising

    that the well-known finite difference has being introduced in order to improve the

    accuracy of isothermal and thermal LBM.

    The first finite difference LBM (FDLBM) was due to Reider and Sterling,10

    and was examined by Cao et al.11 in more detail. FDLBM was further extended

    to curvilinear coordinates with non-uniform grids by Mei and Shyy.12 The study of

    FDLBM is still in progress.13–15 However, there are still no evidence of combination

    of finite difference with thermal lattice Boltzmann.

    The purpose of this paper is to improve the earlier proposed double-distribution

    function thermal lattice Boltzmann method (TLBM).16 Although this model has

    successfully simulated the natural convection problem to a certain degree with

    low computational cost, it is limited for the simulation at low Rayleigh numbers.

    However, for real thermal engineering applications, the value of Rayleigh numbers

    could be achieved up to 106. In this paper, we apply the finite difference technique to

    solve the advection term in the governing equations of double-distribution function

    TLBM. The combination of finite difference with TLBM (FDTLBM) contributes

    in allowing us to increase the accuracy both in time and space where the high

    order accuracy is crucial for the simulation of natural convection at high Rayleigh

    numbers.

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    Finite Difference Thermal Lattice Boltzmann Method 3867

    2. Double-Distribution Function TLBM

    Following the double-distribution function approach proposed by He4 and Nor

    Azwadi and Takahaski,16 the governing equations for these two functions are

    ∂f

    ∂t+ c

    ∂f

    ∂x= − 1

    τv(f − f eq) + Ff , (1)

    ∂g

    ∂t+ c

    ∂g

    ∂x= − 1

    τc(g − geq) , (2)

    where the density distribution function f = f(x, c, t) is used to simulate the

    density and velocity fields, and the internal energy density distribution function

    g = g(x, c, t) is used to simulate the macroscopic temperature field. The macro-

    scopic variables, such as the density ρ, velocity u, and temperature T can be eval-

    uated as the moment to the distribution function

    ρ =

    fdc , ρu =

    cfdc , ρT =

    gdc . (3)

    f eq and geq in Eqs. (1) and (2) are the equilibrium distribution function for density

    and internal energy, respectively, and is given by

    f eq = ρ(1

    2πRT)D/2 exp

    {

    − c2

    2RT

    } [

    1 +c · uRT

    +(c · u)22(RT )2

    − u2

    2RT

    ]

    , (4)

    geq = ρT

    (

    1

    2πRT

    )D/2

    exp

    {

    − c2

    2RT

    }

    [

    1 +c · uRT

    ]

    . (5)

    Equation (5) is obtained by assuming that at low Mach number flow (incompressible

    flow), the higher order of macroscopic velocity and viscous heat dissipation can be

    neglected.6 It has also been proved17 that the above simplification does not alter

    the corresponding macroscopic equation of energy. The only change is the value

    of the constant parameter in the thermal conductivity, which can be absorbed by

    manipulating the parameter τc.

    We have also recently shown that the discretized equilibrium distribution func-

    tion for both density and internal energy density distribution function can be ob-

    tained by applying the Gauss–Hermite quadrature procedure for the calculation of

    f eq and geq velocity moments. As a result, a 2-dimensional 9-velocity, D2Q9 lat-

    tice model, as shown in Fig. 1(left), is obtained, and the corresponding discretized

    equilibrium density distribution function is given by

    f eqi = ρωi

    [

    1 + 3c · uc2

    +9(c · u)2

    2c4− 3u

    2

    2c2

    ]

    , (6)

    where c =√

    3RT and the weights are ω1 = 4/9, ω2 = ω3 = ω4 = ω5 = 1/9,

    and, ω6 = ω7 = ω8 = ω9 = 1/36. While the lattice type for the energy model

    is 2-dimensional 4-velocity, D2Q4 lattice model shown in Fig. 1(right), and the

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    3868 C. S. Nor Azwadi & T. Tanahashi

    5c

    1c

    8c

    6c

    7c

    4c

    3c

    2c

    0c

    1c

    4c

    3c

    2c

    Fig. 1. Lattice structure for D2Q9 (left) and D2Q4 (right).

    corresponding discretized internal energy density equilibrium distribution function

    is given by16

    geq1,2,3,4 =1

    4ρT

    [

    1 +c · uc2

    ]

    . (7)

    Through a multiscaling expansion, the mass and momentum equation can be

    derived from D2Q9 and temperature equation from D2Q4 as below

    ∇ · u = 0 , (8)

    ∂u

    ∂t+ u∇ · u = −1

    ρ∇p + ν∇2u , (9)

    ∂T

    ∂t+ ∇ · (uT ) = χ∇2T . (10)

    The viscosity and thermal diffusivity in these models are related to the time

    relaxations as below

    ν =1

    3τv , (11)

    χ = τc . (12)

    3. Finite Difference Thermal Lattice Boltzmann Method

    (FDTLBM)

    The temporal discretization is obtained using second-order Runge–Kutta (modified)

    Euler method. The time evolution of particle distributions is then derived by

    fn+ 1

    2

    i = fni +

    ∆t

    2

    [

    −ci · ∇fni −1

    τv(fni − f eq,ni )

    ]

    , (13)

    fn+1i = fni + ∆t

    [

    −ci · ∇fn+1

    2

    i −1

    τv

    (

    fn+ 1

    2

    i − feq,n+ 1

    2

    i

    )

    ]

    . (14)

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    Finite Difference Thermal Lattice Boltzmann Method 3869

    The third-order upwind scheme (UTOPIA) was applied to calculate the advec-

    tion term in Eq. (1) as below

    cix∂xfi = cixfi(x + 2∆x, y) − 2fi(x + ∆x, y) + 9fi(x, y)

    6∆x

    + cix−10fi(x − ∆x, y) + 2fi(x − 2∆x, y)

    6∆x, cix > 0 , (15)

    cix∂xfi = cix−fi(x − 2∆x, y) + 2fi(x − ∆x, y) − 9fi(x, y)

    6∆x

    + cix10fi(x + ∆x, y) − 2fi(x + 2∆x, y)

    6∆x, cix < 0 , (16)

    ciy∂yfi = ciyfi(x, y + 2∆y) − 2fi(x, y + ∆y) + 9fi(x, y)

    6∆y

    + ciy−10fi(x, y − ∆y) + 2fi(x, y − 2∆y)

    6∆y, ciy > 0 , (17)

    ciy∂yfi = ciy−fi(x, y − 2∆y) + 2fi(x, y − ∆y) − 9fi(x, y)

    6∆y

    + ciy10fi(x, y + ∆y) − 2fi(x, y + 2∆y)

    6∆y, ciy < 0 . (18)

    The same procedures were carried out for the evolution of temperature equation.

    From this combination, the accuracy of the FDTLBM is second order in time and

    third order in space. The time step used in the computation is varied between 0.1

    and 0.001, depending on the Rayleigh number and mesh size.

    4. Natural Convection in a Square Cavity

    Numerical simulation for the natural convection flow in a square cavity with a

    hot wall on the left side and cool wall on the right side up to Rayleigh number,

    Ra = 106 was carried out to test the effectiveness of the FDTLBM. Figure 2 shows

    a schematic diagram of the setup in the simulation.

    The conventional no-slip boundary conditions1 are imposed on all the walls of

    the cavity. The thermal conditions applied on the left and right walls are T (x =

    0, y) = TH and T (x = L, y) = TC . The top and bottom walls being adiabatic,

    ∂T/∂y = 0.

    The temperature difference between the left and right walls introduces a tem-

    perature gradient in a fluid, and the consequent density difference induces a fluid

    motion, that is, convection.

    In the simulation, the Boussinesq approximation is applied to the buoyancy

    force term.

    ρG = ρβg0(T − Tm)j , (19)

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    3870 C. S. Nor Azwadi & T. Tanahashi

    0,0 =∂

    ∂=

    y

    Tu

    x

    y

    HTT =

    = 0u

    CTT =

    = 0u

    0,0 =∂

    ∂=

    y

    Tu

    0g

    L

    L

    Fig. 2. Schematic geometry for natural convection in a square cavity.

    where β is the thermal expansion coefficient, g0 is the acceleration due to gravity,

    Tm is the average temperature, and j is the vertical direction opposite to that of

    gravity. So the external force in Eq. (1) will be

    Ff = 3G(c − u)f eq . (20)

    The dynamical similarity depends on two dimensionless parameters: the Prandtl

    number, Pr and the Rayleigh number, Ra

    Pr =ν

    χ, Ra =

    g0β∆TL3

    νχ. (21)

    In all simulations, Pr is set to be 0.71 and through the grid dependence study,

    the grid sizes of 101 × 101, 151 × 151, 201 × 201, and 251 × 251 are suitable forRayleigh number 103, 104, 105, and 106, respectively. The convergence criterion for

    all the cases tested is

    Max|(

    (u2 + v2)n+1)

    1

    2 −(

    (u2 + v2)n)

    1

    2 | ≤ 10−7 , (22)

    = Max|T n+1 − T n| ≤ 10−7 , (23)

    where the calculation is carried out over the entire system.

    5. Numerical Results

    Figures 3–6 show the time development of isotherms and their corresponding

    streamlines for all the Rayleigh numbers simulations.

    At the beginning of the simulation for Ra = 103, a vortex appears at the center

    left of the cavity. As time evolves, the vortex is shifted to the center of the cavity.

    The isotherms are almost vertically parallel to the wall, indicating that conduction

    is the dominant heat transfer mechanism. For Ra = 104, a vertically oval-shaped

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    Finite Difference Thermal Lattice Boltzmann Method 3871

    Fig. 3. Time development of isotherms and streamlines for Ra = 103.

    Fig. 4. Time development of isotherms and streamlines for Ra = 104.

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    3872 C. S. Nor Azwadi & T. Tanahashi

    Fig. 5. Time development of isotherms and streamlines for Ra = 105.

    Fig. 6. Time development of isotherms and streamlines for Ra = 106.

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    Finite Difference Thermal Lattice Boltzmann Method 3873

    vortex appears at the center left of the cavity. After that, the vortex is shifted

    to the center of the cavity and its shape changes to horizontal oval due to the

    convection effect. Isotherms start to be horizontally parallel to the wall at the cavity

    center. This indicates that the heat transfer mechanisms are mixed conduction and

    convection.

    For the simulation at Ra = 105, two vortices appear, where one at the top

    left and the other one at the bottom right of the cavity when the system achieved

    equilibrium condition. All isotherms are almost horizontally parallel to the wall,

    indicating that the convection is the main heat transfer mechanism. The vortices

    continue to break up when the Rayleigh number is increased up to 106.

    Figure 7 shows the non-dimensional temperature profile given at the mid-height

    of the cavity for the laminar flow simulations. The profiles show the rapid change in

    the heat transfer mechanisms from conduction to convection. From a 45◦ slope at

    low Rayleigh number, the temperature profiles become horizontal lines in the cavity

    center and all temperature gradients are located in the interior of the boundary

    layer, which has developed near the vertical walls. Near the center of the cavity, the

    curves change slope and there is a vortex corresponding to each change. It can be

    clearly seen that the steep variation of the temperature near the walls is resolved

    quite well.

    Figures 8 and 9 present similar profiles for the horizontal and vertical velocity

    components, respectively. Both figures show a gradually increasing velocity near

    the center, and the development of narrow boundary layers along the walls. The

    peak values of the horizontal and vertical velocities increase due to the intensified

    convective activities with increase in Rayleigh number. The steep rise in the vertical

    velocity gradient at the point on the hot and cold walls also confirms the increased

    310=Ra

    410=Ra

    610=Ra

    510=Ra

    Fig. 7. Non-dimensional temperature profile at the mid-height of the cavity.

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    3874 C. S. Nor Azwadi & T. Tanahashi

    310=Ra

    410=Ra

    610=Ra

    510=Ra

    Fig. 8. Non-dimensional horizontal velocity at the mid-width of the cavity.

    310=Ra

    410=Ra

    610=Ra

    510=Ra

    310=Ra

    410=Ra

    610=Ra

    510=Ra

    Fig. 9. Non-dimensional vertical velocity at the mid-width of the cavity.

    convective activity at higher Rayleigh number values observed in Fig. 9. The changes

    in the velocity direction correspond to slope changes of the temperature profile and

    lead to vortex development.

    In order to validate the present numerical algorithm, the predicted results

    are compared with the results obtained by the Navier–Stokes equation approach.

    Among the characteristic numerical values of the flow, the comparisons concern the

    average Nusselt number at the mid-plane wall, Nuave the maximum value for hor-

    izontal and vertical velocity components, umax and vmax with the positions where

    they occur.

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    Finite Difference Thermal Lattice Boltzmann Method 3875

    Table 1. Comparison between the present results (FDTLBM) and aNavier–Stokes solver.18

    Ra

    103 104 105 106

    umax FDTLBM 3.638 16.127 34.700 65.827

    N–S Solver18 3.634 16.182 34.810 65.330

    y FDTLBM 0.810 0.820 0.855 0.852

    N–S Solver18 0.813 0.823 0.855 0.848

    vmax FDTLBM 3.691 19.584 68.319 221.071

    N–S Solver18 3.679 19.509 68.220 216.750

    x FDTLBM 0.180 0.120 0.065 0.040

    N–S Solver18 0.179 0.120 0.066 0.038

    Nuave FDTLBM 1.117 2.233 4.483 8.723

    N–S Solver18 1.116 2.234 4.510 8.798

    As shown in Table 1, for Ra = 103 and Ra = 104, the present results are in

    close agreement with the Navier–Stokes solution obtained by Davis.18 However,

    small discrepancies can be seen for higher values of Rayleigh numbers, Ra = 105

    and Ra = 106. For all values of Rayleigh number considered in the present analysis,

    the average Nusselt number for the system have been predicted with less than 3%

    error and can be accepted for real engineering applications.

    6. Conclusion

    The natural convection in a differentially heated square cavity has been studied

    using the double-distribution approach thermal lattice Boltzmann method with

    small microvelocity components applied in the internal energy distribution func-

    tion. The evolution of lattice Boltzmann equations have been discretized using the

    third-order accuracy finite difference upwind, UTOPIA scheme. From Figs. 7–9,

    the boundary layers for the velocities and temperature can be observed clearly. As

    expected, the thermal boundary layer is thicker than the velocity boundary layer

    for every Rayleigh number simulations. The flow patterns including the boundary

    layers and vortices can be seen clearly. The results obtained demonstrate that this

    new approach in the double-distribution function thermal lattice Boltzmann model

    is a very efficient procedure to study flow and heat transfer in a differentially heated

    square enclosure.

    Acknowledgments

    The authors wish to acknowledge Universiti Teknologi Malaysia, Keio University

    and the Malaysian government for supporting these research activities.

    References

    1. J. Hardy, Y. Pomeau and D. Pazzis, J. Math. Phys. 14, 1746 (1973).

  • September 3, 2008 10:0 WSPC/140-IJMPB 04861

    3876 C. S. Nor Azwadi & T. Tanahashi

    2. G. McNamara and B. Alder, Phys. A. 194, 218 (1993).3. X. Shan, Phys. Rev. E. 55, 2780 (1997).4. X. He, S. Shan and G. D. Doolen, J. Comp. Phys. 146, 282 (1998).5. H. Chen and C. Teixeira, Comp. Phys. Comm. 129, 21 (2000).6. Y. Peng, C. Shu and Y. T. Chew, Phys. Rev. E 68, 020671 (2003).7. L. S. Luo and X. He, Phys. Rev. E 55, R6333 (1997).8. X. He and L. S. Luo, Phys. Rev. E 56, 6811282 (1997).9. T. Abe, J. Comp. Phys. 131, 241 (1997).

    10. M. B. Reider and J. D. Sterling, Comp. Fluids 24, 459 (1995).11. N. Cao, S. Chen, S. Jin and D. Martinez, Phys. Rev. E 55, R21 (1997).12. Z. R. Mei and W. Shyy, J. Comp. Phys. 143, 426 (1998).13. J. Tolke, M. Krafczvk, M. Schulz, E. Rank and R. Berrios, Int. J. Mod. Phys. C 9,

    1143 (1998).14. G. Hazi, Int. J. Mod. C 13, 67 (1998).15. T. Seta and R. Takahashi, J. Stat. Phys. 107, 557 (2002).16. C. S. Nor Azwadi and T. Takahashi, Int. J. Mod. Phys. B 20, 2437 (2006).17. Z. Guo, Y. Shi and T. S. Zhao, Phys. Rev. E 70, 066310 (2004).18. D. V. Davis, Int. J. Numer. Meth. Fluids 3, 249 (1983).