chew shu niu (ijmpc) 2002

Upload: hamid-mojiry

Post on 04-Jun-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    1/20

    International Journal of Modern Physics C, Vol. 13, No. 6 (2002) 719738c World Scientific Publishing Company

    SIMULATION OF UNSTEADY INCOMPRESSIBLE

    FLOWS BY USING TAYLOR SERIES EXPANSION- AND

    LEAST SQUARE-BASED LATTICE BOLTZMANN METHOD

    Y. T. CHEW, C. SHU, and X. D. NIU

    Department of Mechanical Engineering, National University of Singapore

    10 Kent Ridge Crescent, Singapore 117576

    Received 4 December 2001Revised 31 January 2002

    In this work, an explicit Taylor series expansion- and least square-based lattice Boltz-mann method (LBM) is used to simulate two-dimensional unsteady incompressible vis-cous flows. The new method is based on the standard LBM with introduction of theTaylor series expansion and the least squares approach. The final equation is an explicitform and essentially has no limitation on mesh structure and lattice model. Since theTaylor series expansion is only applied in the spatial direction, the time accuracy of the

    new method is kept the same as the standard LBM, which seems to benefit for unsteadyflow simulation. To validate the new method, two test problems, that is, the vortexshedding behind a circular cylinder at low Reynolds numbers and the oscillating flow ina lid driven cavity, were considered in this work. Numerical results obtained by the newmethod agree very well with available data in the literature.

    Keywords: Lattice Boltzmann equation; explicit method; Taylor series expansion; leastsquare approach; unsteady flow; incompressible.

    1. Introduction

    The study of unsteady fluid systems has been a major interest for fluid dynamicresearchers over almost the last whole century due to its great relevance to en-

    gineering applications in reality. As the need to account for the effect of time-

    dependence presents a considerable difficulty of analysis, numerical simulation plays

    an important role in this field.

    Conventional methods for simulating viscous unsteady flow include, macroscop-

    ically, numerical integration of the NavierStokes equations, and, microscopically,

    molecular dynamics simulation. However, the former has particular difficulty on

    the implementation of complex geometries while the latter is extremely intensive in

    computation. Recently, the lattice Boltzmann method (LBM),1,2

    which originatedfrom the lattice gas cellular automata (LGCA), has provided an alternative ap-

    proach for solving continuum problems on various physical systems. The LBM is

    based on gas-kinetic representations of fluid flow in a strongly reduced particle ve-

    locity space, in which flow is described through the evolution of the discrete particle

    719

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    2/20

    720 Y. T. Chew, C. Shu & X. D. Niu

    distribution functions on uniform lattices. Hydrodynamic variables are computed

    at the lattice nodes as moments of the discrete distribution functions.

    One of the advantages of the LBM is that it can recover the unsteady in-compressible NavierStokes equations with second-order of accuracy in space and

    time at low Knudsen number and low frequency limit through Taylor series and

    ChapmanEnskog expansions.3,4 Thus it is expected that the LBM can simulate

    the unsteady fluid problems in well accuracy. In fact, under the low frequency limit,

    the characteristic timeTof the LBM is in the order of (M a)1t, whereM ais the

    global Mach number, is the Knudsen number and t is the particle streaming

    time. Therefore, the LBM can be applied to the unsteady fluid systems with a

    range of Strouhal numbers ofStr = L/UT = t/TMa O (1) (L and U is the

    characteristic length and velocity, respectively). Many numerical experiments58have confirmed this conclusion.

    The major advantage of LBM is its algebraic form and ease for application.

    However, due to the use of uniform lattice, the broad application of the LBM in

    engineering has been greatly hampered, especially for the problems with curved

    boundaries or with high Reynolds numbers, which need to use a nonuniform mesh

    to obtain high-resolution results in the very thin boundary layer. Theoretically, the

    feature of lattice-uniformity is not necessary to be kept because the distribution

    functions are continuous in physical space. Currently, there are two ways to re-

    move the difficulty of the standard LBM for application to complex problems withnonuniform meshes. One is the so-called interpolation-supplemented LBM (ISLBM)

    proposed by He and his colleagues.79 They successfully applied this approach to

    simulate flows past an impulsively started cylinder. The other is based on the so-

    lution of a differential lattice Boltzmann equation (LBE). For complex problems,

    the differential LBE can be solved by the finite difference (FDLBE) method with

    the help of coordinate transformation10 or by the finite volume method.11,12 These

    methods have been successfully applied to solve quite a number of complex prob-

    lems. However, the ISLBE has an extra computational effort for interpolation at

    every time step, and it also has a strict restriction on the selection of interpo-lation points. For the FDLBE and FVLBE methods, one need to select efficient

    approaches such as upwind schemes to do numerical discretization in order to get

    the stable solution. As a consequence, the computational efficiency greatly depends

    on the selected numerical scheme. In addition, the time accuracy of FDLBE and

    FVLBE methods is reduced due to truncation error of discretization in time as

    compared to the standard LBM.

    In the present work, an explicit Taylor series expansion- and least square-based

    LBM is proposed. The new method is based on the standard LBM with introduction

    of Taylor series expansion in the spatial direction and least squares optimization.The final form of the method is an algebraic formulation, in which the coefficients

    depend only on the coordinates of mesh points and lattice velocity, and are com-

    puted in advance. The new method can be easily applied to different other lattice

    models. Since the Taylor series expansion is applied only in the spatial direction, the

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    3/20

    Simulation of Unsteady Incompressible Flows 721

    time accuracy of the new method is kept the same as the standard LBE. Thus, it

    is expected that the new method can well simulate the unsteady flow problems. To

    validate this, two unsteady flows problems are considered in this work. One is thevortex shedding flow behind a circular cylinder. The investigation of vortex shed-

    ding in the wake of a circular cylinder at low Reynolds number is very important

    because it can provide some basic insights into the vortex shedding mechanism.

    The other case is the oscillating flow in a lid driven cavity. This problem is often

    chosen to test new numerical schemes for simulation of unsteady flows. There are

    many publications8,1322 on these two cases in the literature. So, it is ideal to use

    them as unsteady benchmarks to verify the present method.

    2. Taylor Series Expansion- and Least Squares-Based LBM

    The method developed in this work is based on the fact that the distribution func-

    tion is a continuous function in physical space and can be well defined in any mesh

    system. Let us start with the standard LBM. The two dimensional, standard LBE

    with BGK approximation can be written as:

    f(x+ext,y+eyt,t+t)

    =f(x,y,t) +feq (x,y,t) f(x,y,t)

    , = 0, 1, . . . , M , (1)

    where is the single relaxation time; f is the density distribution function along

    the direction; feq is its corresponding equilibrium state, which depends on the

    local macroscopic variables such as a density and velocity U(u, v); t is the time

    step and e(ex, ey) is the particle velocity in the direction; M is the number

    of discrete particle velocities. Obviously, the standard LBE consists of two steps:

    collision and streaming. The macroscopic density and momentum density Uare

    defined as:

    =

    M=0

    f, U=

    M=0

    fe. (2)

    Suppose that a particle is initially at the grid point (x,y,t). Along the direc-

    tion, this particle will stream to the position (x+ ext,y+eyt,t+ t). For a

    uniform lattice, x = ext, y = eyt. So, (x+ ext, y+ eyt) is on the grid

    point. In other words, Eq. (1) can be used to update the density distribution func-

    tions exactly at the grid points. However, for a nonuniform grid, (x+ext, y+eyt)

    is usually not at the grid point (x+ x, y+ y). In the numerical simulation, we

    are interested only in the density distribution function at the mesh point for all

    the time levels. So, the macroscopic properties such as the density, flow velocitycan be evaluated at every mesh point. To get the density distribution function at

    the mesh point (x+ x, y +y) and the time level t+ t, we need to apply the

    Taylor series expansion or other interpolation techniques such as the one used by

    He et al.79 In this work, the Taylor series expansion is used. Note that the time

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    4/20

    722 Y. T. Chew, C. Shu & X. D. Niu

    P

    A

    B

    C

    D

    E

    A'

    B'

    C'

    D'

    P'

    E'

    Fig. 1. Configuration of particle movement along the direction.

    level for the position (x + ext, y + eyt) and the grid point (x + x,y + y) is the

    same, that is, t+t. So, the expansion in the time direction is not necessary. As

    shown in Fig. 1, we let point A represent the position (xA, yA) point A represent

    the position (xA+ ext, yA+ eyt), and point Prepresent the position (xP, yP).

    Using Eq. (1), we can get the density distribution function at the position A

    f(A, t+t) =f(A, t) +

    feq (A, t) f(A, t)

    . (3)

    For the general case, A

    may not coincide with the mesh point P. In this case, weneed to obtain the density distribution function at the mesh point P. This can be

    done by applying the Taylor series expansion in the spatial direction only. With

    Taylor series expansion, f(A, t+t) can be approximated by the corresponding

    function and its derivatives at the mesh point P as:

    f(A, t+t) =f(P, t+t) + xA

    f(P, t+t)

    x + yA

    f(P, t+t)

    y

    +1

    2(xA)

    2 2f(P, t+t)

    x2 +

    1

    2(yA)

    2 2f(P, t+t)

    y2

    + xAyA2f(P, t+t)

    xy +O[(xA)

    3, (yA)3] , (4)

    where xA = xA + ext xp, yA = yA +eyt yp. Note that the above

    approximation has a truncation error of the third order. Substituting Eq. (4) into

    Eq. (3) gives

    f(P, t+t) + xAf(P, t+t)

    x

    + yAf(P, t+t)

    y

    +1

    2(xA)

    2 2f(P, t+t)

    x2 +

    1

    2(yA)

    2 2f(P, t+t)

    y2

    + xAyA2f(P, t+t)

    xy =f(A, t) +

    feq (A, t) f(A, t)

    . (5)

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    5/20

    Simulation of Unsteady Incompressible Flows 723

    It is indicated that Eq. (5) is a differential equation, which involves only two

    mesh pointsAand P. Solving Eq. (5) can provide the density distribution functions

    at all the mesh points. Equation (5) can be considered as a new version of differentialLBE, which can give very accurate numerical results. In this work, we go further

    to develop a new solution procedure. In fact, our new development is inspired from

    the RungeKutta method. As we know, the RungeKutta method is developed to

    improve the Taylor series method in the solution of ordinary differential equations

    (ODEs). Like Eq. (5), Taylor series method involves evaluation of different orders of

    derivatives to update the functional value at the next time level. For a complicated

    expression of given ODEs, this application is very difficult. To improve the Taylor

    series method, the RungeKutta method evaluates the functional values at some

    intermediate points and then combines them (through the Taylor series expansion)to form a scheme with the same order of accuracy. With this idea in mind, we look

    at Eq. (5). We know that at the time level t + t, the density distribution function

    and its derivatives at mesh pointPare all unknowns. So, Eq. (5) has six unknowns

    in total. To solve for the six unknowns, we need six equations. However, Eq. (5)

    just provides one equation. We need additional five equations to close the system.

    As shown in Fig. 1, we can see that along the direction, the particles at five mesh

    points P, B, C, D, E at the time level t will stream to the new positions P, B,

    C,D, E at the time level t + t. The density distribution functions at these new

    positions can be computed through Eq. (1), which are given below

    f(P, t+t) =f(P, t) +

    feq (P, t) f(P, t)

    , (6)

    f(B, t+t) =f(B, t) +

    feq (B, t) f(B, t)

    , (7)

    f(C, t+t) =f(C, t) +

    feq (C, t) f(C, t)

    , (8)

    f(D

    , t+t) =f(D, t) +

    feq (D, t) f(D, t)

    , (9)

    f(E, t+t) =f(E, t) +

    feq (E, t) f(E, t)

    . (10)

    Using Taylor series expansion, f(P, t + t), f(B, t + t), f(C, t + t),

    f(D, t+t), f(E, t+t) in above equations can be approximated by the func-

    tion and its derivatives at the mesh point P. As a result, Eqs. (6)(10) can be

    reduced to:

    f(P, t+t) + xPf(P, t+t)

    x

    + yPf(P, t+t)

    y

    +1

    2(xP)

    2 2f(P, t+t)

    x2 +

    1

    2(yP)

    2 2f(P, t+t)

    y2

    + xPyP2f(P, t+t)

    xy =f(P, t) +

    feq (P, t) f(P, t)

    , (11)

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    6/20

    724 Y. T. Chew, C. Shu & X. D. Niu

    f(P, t+t) + xBf(P, t+t)

    x + yB

    f(P, t+t)

    y

    +1

    2(xB)

    2 2f(P, t+t)

    x2 +

    1

    2(yB)

    2 2f(P, t+t)

    y2

    + xByB2f(P, t+t)

    xy =f(B, t) +

    feq (B, t) f(B, t)

    , (12)

    f(P, t+t) + xCf(P, t+t)

    x + yC

    f(P, t+t)

    y

    +

    1

    2(xC)2

    2f(P, t+t)

    x2 +

    1

    2(yC)2

    2f(P, t+t)

    y2

    + xCyC2f(P, t+t)

    xy =f(C, t) +

    feq (C, t) f(C, t)

    , (13)

    f(P, t+t) + xDf(P, t+t)

    x + yD

    f(P, t+t)

    y

    +1

    2(xD)

    2 2f(P, t+t)

    x2 +

    1

    2(yD)

    2 2f(P, t+t)

    y2

    + xDyD2f(P, t+t)

    xy =f(D, t) +

    feq (D, t) f(D, t)

    , (14)

    f(P, t+t) + xEf(P, t+t)

    x + yE

    f(P, t+t)

    y

    +1

    2(xE)

    2 2f(P, t+t)

    x2 +

    1

    2(yE)

    2 2f(P, t+t)

    y2

    + xEyE2f(P, t+t)

    xy =f(E, t) +

    feq

    (E, t) f(E, t)

    , (15)

    where

    xP =ext , yP =eyt ,

    xB =xB+ ext xP, yB =yB+ eyt yP,

    xC =xC+ext xP, yC =yC+eyt yP,

    xD =xD+ext xP , yD =yD+eyt yP,

    xE

    =xE

    +ex

    t xP

    , yE

    =yE

    + ey

    t yP

    .

    Equations (5), (11)(15) form a system to solve for six unknowns. Now, we

    define

    gi=f(xi, yi, t) +feq (xi, yi, t) f(xi, yi, t)

    , (16)

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    7/20

    Simulation of Unsteady Incompressible Flows 725

    {s}T =

    1, xi, yi,

    (xi)2

    2 ,

    (yi)2

    2 , xiyi

    , (17)

    {V} =

    f,

    fx

    , f

    y ,

    2fx2

    ,2f

    2y ,

    2fxy

    T, (18)

    wheregi is the post-collision state of the distribution function at the ith point and

    the time level t, {si}T is a vector with six elements formed by the coordinates of

    mesh points, {V} is the vector of unknowns at the mesh point P, which also has

    six elements. Our target is to find its first element V1 =f(P, t+t). With above

    definitions, Eqs. (5), (11)(15) can be written as:

    gi= {si}T{V} =

    6j=1

    si,jVj, i= P, A, B,C,D, E , (19)

    where si,j is the jth element of the vector {si}T and Vj is the jth element of the

    vector{V}.

    Equation system (19) can be put into the following matrix form

    [s]{V} = {g} , (20)

    where

    {g} = {gP, gA, gB, gC, gD, gE}T

    ,

    [S] = [si,j ] =

    {sP}T

    {sA}T

    {sB}T

    {sC}T

    {sD}T

    {sE}T

    =

    1 xP yP(xP)2

    2

    (yP)2

    2 xPyP

    1 xA yA(xA)

    2

    2

    (yA)2

    2 xAyA

    1 xB yB(xB)

    2

    2

    (yB)2

    2 xByB

    1 xC

    yC

    (xC)2

    2

    (yC)2

    2 x

    Cy

    C

    1 xD yD(xD)

    2

    2

    (yD)2

    2 xDyD

    1 xE yE(xE)

    2

    2

    (yE)2

    2 xEyE

    .

    Note that when lattice velocity is specified, the matrix [S] depends only on

    the coordinates of mesh points, which can be computed once and stored for the

    application of Eq. (20) at all time levels. In practical applications, it was found

    that matrix [S] might be singular or ill conditioned. To overcome this difficulty andmake the method be more general, we propose the following least squares-based

    LBM.

    Equation (19) has six unknowns (elements of the vector {V}). If Eq. (19) is

    applied at more than six mesh points, then the system is over-determined. For this

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    8/20

    726 Y. T. Chew, C. Shu & X. D. Niu

    case, the unknown vector can be decided from the least square method. For simplic-

    ity, let the mesh pointPbe represented by the indexi= 0, and its adjacent points

    be represented by index i = 1, 2, . . . , N , where N is the number of neighboringpoints around P and it should be larger than 5. At each point, we can define an

    error in terms of Eq. (19), that is,

    erri =gi 6

    j=1

    si,jVj, i= 0, 1, 2, . . . , N . (21)

    The square sum of all the errors is defined as:

    E=

    N

    i=0 err

    2

    i =

    N

    i=0

    gi

    6

    j=1 si,jVj

    2

    . (22)

    To minimize the error E, we need to set E/Vk = 0, k = 1, 2, . . . , 6, which

    leads to:

    [S]T[S]{V} = [S]T{g} , (23)

    where [S] is a (N+ 1) 6 dimensional matrix, which is given as:

    [S] =

    1 x0 y0(x0)2

    2

    (y0)2

    2 x0y0

    1 x1 y1(x1)2

    2

    (y1)2

    2 x1y1

    1 xN yN(xN)2

    2

    (yN)2

    2 xNyN

    (N+1)6

    and {g} = {g0, g1, . . . , gN}T

    .The x and y values in the matrix [S] are given as:

    x0 =ext, y0=eyt , (24a)

    xi =xi+ext x0, yi =yi+eyt y0, for i= 1, 2, . . . , N . (24b)

    Clearly, when the coordinates of mesh points are given, and the particle velocity

    and time step size are specified, the matrix [S] is determined. Then from Eq. (23),

    we obtain

    {V} = ([S]

    T

    [S])

    1

    [S]

    T

    {g} = [A]{g} . (25)Note that [A] is a 6 (N+ 1) dimensional matrix. From Eq. (25), we can have

    f(x0, y0, t+t) =V1

    N+1k=1

    a1,kgk1, (26)

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    9/20

    Simulation of Unsteady Incompressible Flows 727

    01

    23

    4

    5 6

    Fig. 2. Schematic plot of D2Q7 model on a solid boundary (thick black line).

    where a1,k are the elements of the first row of the matrix [A], which are pre-

    computed before the LBM is applied. Therefore, little computational effort is intro-

    duced as compared with the standard LBE. Note that the function g is evaluated

    at time level t. So, Eq. (26) is actually an explicit form to update the distribution

    function at time level t+ t for any mesh point. In the above process, there is

    no requirement for the selection of neighboring points. In other words, Eq. (26) is

    nothing to do with the mesh structure. It needs only to know the coordinates of

    the mesh points. Thus, we can say that Eq. (26) is basically a meshless form.

    It can be seen that Eq. (26) is applied along the direction. Here can be

    any direction. This implies that Eq. (26) can be uniformly applied to the different

    lattice models. In this work, we use the D2Q7 model. The configuration of this

    model is shown in Fig. 2. The discrete velocity of this model is defined as:

    e=

    (0, 0) , = 0 ,

    cos

    ( 1)

    3 , sin

    ( 1)

    3 c , = 1, 2, . . . , 6 .

    (27)

    The parameter c is the particle streaming speed. The fluid kinetic viscosity isgiven by:

    = 2 1

    8 c2t, (28)

    and the equilibrium density distribution feq is chosen to be

    feq =

    1

    2+

    1

    6

    2

    e U

    c2 + 4

    e U

    c2

    2

    U2

    c2

    . (29)

    The speed of sound of this model is cs = c/2, and the equation of state isP = c2s for an ideal gas. Although the proposed method has meshless feature,

    it is recommended to use a structured grid. This is because in our method, only

    the coordinates of mesh points are involved. When a structured grid is used, it is

    easy for us to define the coordinates of mesh points. In our application, we use a

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    10/20

    728 Y. T. Chew, C. Shu & X. D. Niu

    1 1,

    i j+ +1

    ,

    i j+1 1

    ,

    i j +

    1,i j

    1 1,i j 1,i j 1 1,i j+

    1,i j+,i j

    Fig. 3. Schematic plot of neighboring point distribution around the point (i, j).

    structured grid, and takeNas 8 for convenience. As shown in Fig. 3, for an internal

    mesh point (i, j) [noted as 0 in Eq. (26)], the eight neighboring points are taken

    as (i 1, j 1); (i 1, j); (i 1, j+ 1); (i, j 1); (i, j+ 1); (i + 1, j 1); (i + 1, j);

    (i + 1, j+ 1). Therefore, at each mesh point, we need only to store nine coefficients

    a1,k, k = 1, 2, . . . , 9 before Eq. (26) is applied. Note that the configuration of nine

    mesh points as shown in Fig. 3 is applied in all lattice directions ( = 1, 2, . . . , 6).

    Implementation of boundary conditions is an essiential issue in LBM. In this

    work, we found that a complete half-way wall bounceback condition6 is the most

    simple and efficient method in implementing the boundary condition on the soild

    wall, where the nonslip condition holds. The complete half-way wall bounceback

    condition, which originated from LGCA, assigns each f the value of the f in its

    opposite direction with no relaxation on the bounceback points. The treament is

    independent of the direction, which gives us more conveniences in treating compli-

    cated boundary problems. The complete half-way wall bounceback condition has

    second order of accuaracy because macroscopic quantities such as stress force is

    evaluated on the half-way wall between the bounceback row and the first flow row.

    As shown in Fig. 2, the flow field is below the solid boundary represented by the

    black thick line. For the D2Q7 model, at a boundary point, f4, f5 and f6 point to

    the flow field from the wall, which will be determined from the boundary condition.

    f1, f2, f3 are computed by streaming from points inside the flow field. Note that

    whenf1, f2, f3 are computed by Eq. (26), all the neighboring points involved must

    be inside the flow field. In other words, the configuration as shown in Fig. 3 cannot

    be used for this case. We need to select the neighboring points from one side(bottom

    part of the wall as shown in Fig. 2). Using the half-way wall bounceback condition,

    f4, f5 and f6 are evaluated as:

    f4=f1, f5 =f2 , f6=f3. (30)

    3. Results and Discussion

    3.1. Vortex shedding behind a circular cylinder

    As we know, a periodic vortex shedding will occur for flow behind a circular

    cylinder when Reynolds number is above the critical value of 49, and the flow

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    11/20

    Simulation of Unsteady Incompressible Flows 729

    field is two-dimensional when Reynolds number is below the second critical value

    around 149. Above this value, three-dimensional structure of the flow field be-

    comes essential.13,18 In this part, our study focuses only on two-dimensional vor-tex shedding behind a circular cylinder. For this reason, the Reynolds numbers

    (Re =U D/), based on the upstream velocity Uand the diameter of the cylinder

    D, are chosen to be 50, 100 and 150.

    In the flow domain, the following coordinate stretching is used in the radial

    direction

    r= r0+ (r r0)

    1

    1

    tan1[(1 )tan()]

    , (31)

    where r0 = 1 is the cylinder radius, r is the outer boundary, is the parameter

    to control the coordinate stretching and = (j 1)/(jmax 1) with j representingthe index of a mesh point in the radial direction. Uniform grid in the direction is

    adopted.

    Three boundary conditions are required in our simulation for this problem. One

    is at the cylinder surface, where a complete half-way wall bounce back condition

    described in the previous section is used. Another is on the cut line in the wake,

    where the periodic boundary condition is imposed. The third is at the far field

    boundary r where the free stream flow is taken and the density distribution

    function is always set to its equilibrium state. A sketch of the problem is shown in

    Fig. 4. Initially, an asymmetrical flow field

    u= Ur0y

    r2 , v= U

    r0x

    r2 , (32)

    is imposed to serve as an artificial initiator for the vortex shedding process. The

    far field velocityUand pressure are set to 0.15 and 1.0, respectively. The mesh size

    of 241 241 is used for all of the above three Reynolds numbers (typical mesh is

    shown in Fig. 5). The far field boundary is set at 25.5 diameters away from the

    center of the cylinder and the time step, in units ofD/(2U), is equal to 0.00375,

    which corresponds to = 0.72. Our experience showed that this is sufficient to

    capture all the details of the flow at these three Reynolds numbers.

    Periodic BC

    y

    x

    q

    R

    U

    Fig. 4. A sketch of the flow past a circular cylinder.

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    12/20

    730 Y. T. Chew, C. Shu & X. D. Niu

    Fig. 5. Computational mesh for flow around a circular cylinder.

    The most attractive feature of the vortex shedding behind a circular cylinder is

    the periodic variation of the flow field. This periodicity can be illustrated by the

    time evolution of two characteristic parameters, the drag coefficient CD and the lift

    coefficientCL. They are defined as:

    CD = F x

    (1/2)U2D, CL =

    F y(1/2)U2D

    , (33)

    whereF =

    [pI+ (U + U)] ndl. nis the normal vector of the cylinder sur-

    face. The vortex shedding frequency f, which is expressed by the Strouhal number

    ofS t= f D/U, can be obtained by measuring the final period of the lift coefficient.

    Time

    DragCoefficient

    0 50 100 150 2000

    0.25

    0.5

    0.75

    1

    1.25

    1.5

    1.75

    2

    2.25

    2.5

    2.75

    3

    (a) Whole history

    Time

    DragCoefficient

    150 160 170 180 190 200

    1.25

    1.3

    1.35

    1.4

    1.45

    1.5

    (b) Close-up

    Fig. 6. Time evolution of drag coefficients for different Reynolds numbers ( Re = 50;Re = 100; Re = 150).

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    13/20

    Simulation of Unsteady Incompressible Flows 731

    Time

    LiftCoefficient

    0 50 100 150 200-1

    -0

    .75

    -0

    .5

    -0

    .25

    0

    0.25

    0.5

    0.75

    1

    (a) Whole history

    Time

    LiftCoefficient

    150 160 170 180 190 200-1

    -0.75

    -0.5

    -0.25

    0

    0.25

    0.5

    0.75

    1

    (b) Close-up

    Fig. 7. Time evolution of lift coefficients for different Reynolds numbers ( Re = 50;Re = 100; Re = 150).

    Figures 6 and 7 show the time evolution of the drag and lift coefficients for

    different Reynolds numbers. The initially irregular variations of these coefficients

    can be attributed to the initial disturbance. After a certain time, they gradually

    evolve to periodic oscillations, which become stronger with increase of the Reynolds

    number. The lift coefficient has a stronger oscillation than the drag coefficient. From

    the close-up of these two figures, one can observe that the variation of the drag

    coefficient is two times faster than that of lift coefficient. The reason may be due to

    the fact that the drag coefficient is mainly affected by the vortex shedding processes

    from both sides of the cylinder. This observation has been justified by the study of

    Braza et al.19

    Table 1. Comparisons of average and oscillatory drag and lift

    coefficients and Strouhal numbers with previous studies.

    Reynolds number 50 100 150

    Williamson14 0.123 0.164 0.183St He et al.8 0.121 0.161 0.179

    Present 0.123 0.164 0.176

    Tritton20 1.450 1.350 1.330CD He et al.

    8 1.394 1.250 1.261Present 1.461 1.3668 1.314

    Braza et al.19 0.03

    CD He et al.8

    0.002 0.018 0.048Present 0.002 0.0242 0.5

    Braza et al.19 0.60 CL He et al.

    8 0.11 0.64 0.98Present 0.18 0.75 1.08

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    14/20

    732 Y. T. Chew, C. Shu & X. D. Niu

    Table 1 lists the detailed comparison of the average values and oscillations (peak

    to peak) of the drag and lift coefficients, and the calculated Strouhal number St

    with those of the previous studies.8,14,19,20 In Table 1,CD is the mean value of thedrag coefficient, while CD and CL are oscillations of drag and lift coefficients

    from peak to peak. The results of He and Doolen8 were obtained by the LBM with

    interpolation technique. The results of Tritton20 were experimental data. Other

    data shown in Table 1 were the results of NavierStokes equations. Obviously, the

    present results agree very well with those published previously. As compared to

    the experimental data of Tritton,20 it seems that the present results have a better

    accuracy than those of He and Doolen.8

    Since the flow patterns are similar for all the simulated Reynolds numbers, only

    the flow patterns at Re = 100 are presented in the paper. Figure 8 shows the se-quence of vortex shedding over a complete cycle using a sequence of instantaneous

    streamlines and vorticity contours separated by intervals of T /8, where T is the

    period of the shedding cycle. From the evolution of the streamlines (left of Fig. 8),

    one can observe that a large vortex sheds from the top while a small vortex appears

    at the bottom of the cylinder initially. In the meantime, fluid below the cylinder is

    drawn down into the large recirculation region [Fig. 8(a)]. The small vortex form-

    ing at the bottom of the cylinder grows and the large vortex travels to downstream

    (a)t= T /8

    (b) t= T /4

    (c) t= 3T /8

    Fig. 8. Streamlines (left) and vorticity contours (right) near the wake separated by an intervalofT /8.

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    15/20

    Simulation of Unsteady Incompressible Flows 733

    (d)t= T /2

    (e) t= 5T /8

    (f) t= 3T /4

    (g) t= 7T /8

    (h) t= T

    Fig. 8. (Continued)

    while its strength is reduced gradually [Figs. 8(b)8(d)]. When the strength of the

    small vortex grows to its maximum, it breaks off. After that, another shedding pro-

    cess from the bottom of the cylinder is repeated [Figs. 8(e)8(h)]. The mechanismof the vortex shedding is reflected also on the instantaneous vorticity contours. As

    shown in Fig. 8 (right), the alternative change of the strength of a pair of asymmet-

    ric secondary vortices is just the initiator of the vortex shedding from the cylinder.

    The flow patterns in Fig. 8 match well with the numerical simulations by Eaton21

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    16/20

    734 Y. T. Chew, C. Shu & X. D. Niu

    and experimental observation by Perry et al.22 The centers (points surrounded by

    closed streamlines), saddles (points where a streamline crosses itself), separatrices

    (streamlines containing a saddle point) and instant alleyways between two sep-aratrices described by them are observed also in our simulation. However, we did

    not observe the coexistence of centers and saddles of two shed vortices suggested

    by Perry et al.22 Our simulation result is consistent with the simulation results of

    He and Doolen8 and Eaton.21

    3.2. Oscillating flow in a lid driven cavity atRe = 400

    Another test case is the oscillating flow in a square cavity. A periodic velocity

    waveform is imposed on the cavity lid and the time evolution observed in the flowis compared with the numerical results of Soh and Goodrich.16 The periodic lid

    velocity is given by a sinusoidal waveform:

    u(t) =Ucos(t) , (34)

    where U is the maximum lid velocity during the cycle, is the frequency of the

    oscillation andt is the time. The period of oscillation,T, is related to the frequency

    byT = 2/.

    Our simulation is performed with Re = 400 (Re = UL/, where L is the char-

    acteristic length set as the value of top wall), frequency of = 1 and U = 0.15.

    Initially, a periodic velocity given by Eq. (34) is imposed. A nonuniform grid of

    97 97 with more densely distribution near the boundaries is used, and as a con-

    sequence, the time step is set as, in the unit ofL/U, 4.5 104. The oscillatory

    flow reaches the stable periodic state when the velocity components u and v of each

    point in the domain at two consequent flow cycles are with a small tolerance of

    = 105.

    (a) Initial evolution (b) Later stage of evolution

    Fig. 9. Time evolution of drag on the cavity lid.

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    17/20

    Simulation of Unsteady Incompressible Flows 735

    It was found that the flow reaches the stable periodic state after seven cycles.

    Figure 9 shows the time evolution of the viscous drag on the lid, which is estimated

    using the same formula as used by Soh and Goodrich.16 It can be observed thatthe drag settles down to be periodic very quickly, much quicker than the entire flow

    field. The value of maximum drag obtained by the present method agrees very well

    with that predicted by Soh and Goodrich.16

    The instantaneous streamlines obtained by the present method match very well

    with those of Soh and Goodrich,16 which are shown in Figs. 10 and 11 at t =

    (5 +)T where = 0.2, 0.3, 0.35, 0.4, 0.45, 0.5, 0.7, 0.8, 0.85, 0.9, 0.95 and 1.0,

    respectively. Figure 10 shows the streamlines in the first half of the cycle while

    Fig. 11 displays the streamlines in the second half of the cycle. As time advances,

    the direction of the lid movement and the center of the vortex change. The lidvelocity passes through zero at t = T /4, after which it reverses its direction. As a

    result, at t = 0.3T, a counter rotating vorticity is formed in the flow field at the

    left corner of the cavity. As the magnitude of the velocity increases in the negative

    x direction, the size of the second vortex created in the upper left corner of the

    cavity also increases. At the same time, the primary vortex continuously shrinks

    until t = T /2. At this point, the velocity reaches its maximum in the negative x

    direction and the second vortex, which has formed in the left corner of the cavity,

    (a) t= 0.2T (b) t= 0.3T (c) t= 0.35T

    (d) t= 0.4T (e) t= 0.45T (f) t= 0.5T

    Fig. 10. Instantaneous streamlines in the first half of a cycle obtained by the present method.

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    18/20

    736 Y. T. Chew, C. Shu & X. D. Niu

    (a)t= 0.7T (b) t= 0.8T (c)t= 0.85T

    (d) t= 0.9T (e) t= 0.95T (f) t= T

    Fig. 11. Instantaneous streamlines in the second half of a cycle obtained by the present method.

    attains its maximum size and occupies the entire domain. After this point, the

    streamlines at each time step are the mirror images of the streamlines at the time

    t T /2. This finding can be obviously observed by comparing the results of Fig. 10

    with those of Fig. 11. The same finding has been shown in the results of Soh

    and Goodrich.16

    3.3. Effect of grid size

    To show that the present results are accurate, we performed numerical simulations

    for the vortex shedding flow at Re = 100 and the oscillating flow in a lid driven

    cavity by using two different mesh sizes. For the vortex shedding case, two mesh

    sizes of 201 201 and 241 241 are used and the time steps are correspondingly

    set to be, in units ofD/(2U), 0.0045 and 0.00375. For the oscillating flow in the

    lid driven cavity, the mesh sizes of 81 81 and 97 97 with the same time step

    are used. The computed Strouhal number St, average value of drag coefficient,

    oscillations (peak to peek) of drag and lift coefficients for the vortex shedding case,and the oscillation of drag and phase shift for the oscillating flow case are listed in

    Table 2. It was found from Table 2 that all discrepancies caused by different mesh

    sizes are less than 5%. This implies that the mesh sizes of 241 241 for the vortex

    shedding flow and 97 97 for the oscillating flow used in this work are fine enough

    to provide accurate numerical results.

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    19/20

    Simulation of Unsteady Incompressible Flows 737

    Table 2. Effect of grid size.

    Vortex shedding behind the cylinder at Re = 100

    Grid St CD CD CL

    201 201 0.156 1.3363 0.024 0.71241 241 0.164 1.3668 0.0242 0.75

    Oscillating flow in the lid driven cavity at Re = 400

    Grid Oscillation of drag Phase shift

    81 81 58.968 16.0897 97 59.713 15.96

    4. Conclusions

    An explicit Taylor series expansion- and least square-based lattice Boltzmann

    method is presented, and applied to simulate the two-dimensional unsteady vor-

    tex shedding flow behind a circular cylinder and the oscillating flow in the lid

    driven cavity. The dynamic parameters and flow patterns obtained by the present

    method are in good agreement with those of previous numerical and experimental

    studies. The success of this study shows that the present method can be used as

    a versatile CFD tool to simulate real unsteady flows. The major advantage of the

    present scheme is that it still keeps the local and explicit features of the standardlattice Boltzmann method. Since the method has a meshless feature and no coor-

    dinate transformation is involved, it is very convenient for the present method to

    be applied to the flow problems with complex geometry.

    References

    1. G. R. McNamara and G. Zanetti, Phys. Rev. Lett. 61, 2332 (1988).2. F. J. Higuera, S. Succi, and R. Benzi, Europhys. Lett. 9, 345 (1989).3. H. Chen, S. Chen, and W. H. Matthaeus,Phys. Rev. A 45, R5339 (1991).

    4. Y. H. Qian, D. dHumieres, and P. Lallemand, Europhys. Lett. 17, 479 (1992).5. S. Chen and G. D. Doolen, Ann. Rev. Fluid Mech. 30, 329 (1998).6. Q. Zou and X. He, Phys. Fluids9, 1591 (1997).7. X. He and G. D. Doolen,J. Comp. Phys. 134, 306 (1997).8. X. He and G. D. Doolen,Phys. Rev. E56, 434 (1997).9. X. He, L.-S. Luo, and M. Dembo,J. Comp. Phys. 129, 357 (1996).

    10. R. Mei and W. Shyy, J. Comp. Phys. 143, 426 (1998).11. H. Chen,Phys. Rev. E58, 3955 (1998).12. G. Peng, H. Xi, C. Duncan, and S. H. Chou,Phys. Rev. E59, 4675 (1999).13. E. Berger and R. Wille, Ann. Rev. Fluid Mech. 4, 313 (1972).

    14. C. H. K. Williamson,Phys. Fluids31

    , 2742 (1988).15. M. Hammache and M. Gharib,Phys. Fluids A 1, 1611 (1989).16. W. Y. Soh and J. W. Goodrich, J. Comp. Phys. 79, 113 (1988).17. A. L. Gaitinde, Int. J. Numer. Meth. Eng. 41, 1153 (1998).18. C. H. K. Williamson, Ann. Rev. Fluid Mech. 23, 477 (1996).19. M. Braza, P. Chassaing, and M. H. Ha, J. Fluid Mech. 165, 79 (1986).

  • 8/14/2019 Chew Shu Niu (Ijmpc) 2002

    20/20

    738 Y. T. Chew, C. Shu & X. D. Niu

    20. D. J. Tritton,J. Fluid Mech. 6, 547 (1959).21. B. E. Eaton,J. Fluid Mech. 6, 547 (1959).

    22. A. E. Perry, M. S. Chong, and T. T. Lim, J. Fluid Mech. 116

    , 77 (1982).