Appl. Math. Mech. -Engl. Ed.   2014, Vol. 35 Issue (12): 1479-1494     PDF       
http://dx.doi.org/10.1007/s10483-014-1885-6
Shanghai University
0

Article Information

Zheng TANG, Nan-sheng LIU, Yu-hong DONG. 2014.
Lattice Boltzmann simulations of turbulent shear flow between parallel porous walls
Appl. Math. Mech. -Engl. Ed., 35(12): 1479-1494
http://dx.doi.org/10.1007/s10483-014-1885-6

Article History

Received 2014-1-15;
in final form 2014-5-4
Lattice Boltzmann simulations of turbulent shear flow between parallel porous walls
Zheng TANG1, Nan-sheng LIU3, Yu-hong DONG1,2        
1. Shanghai Institute of Applied Mathematics and Mechanics, Shanghai University, Shanghai 200072, P. R. China;
2. Shanghai Key Laboratory of Mechanics in Energy Engineering, Shanghai University, Shanghai 200072, P. R. China;
3. Department of Modern Mechanics, University of Science and Technology of China, Hefei 230026, P. R. China
ABSTRACT:The effects of two parallel porous walls are investigated, consisting of the Darcy number and the porosity of a porous medium, on the behavior of turbulent shear flows as well as skin-friction drag. The turbulent channel flow with a porous surface is directly simulated by the lattice Boltzmann method (LBM). The Darcy-Brinkman-Forcheimer (DBF) acting force term is added in the lattice Boltzmann equation to simulate the turbulent flow bounded by porous walls. It is found that there are two opposite trends (enhancement or reduction) for the porous medium to modify the intensities of the velocity fluctuations and the Reynolds stresses in the near wall region. The parametric study shows that flow modification depends on the Darcy number and the porosity of the porous medium. The results show that, with respect to the conventional impermeable wall, the degree of turbulence modification does not depend on any simple set of parameters obviously. Moreover, the drag in porous wall-bounded turbulent flow decreases if the Darcy number is smaller than the order of O(10-4) and the porosity of porous walls is up to 0.4.
Keywordslattice Boltzmann method (LBM)     porous material     wall-bounded turbulence     Darcy number     drag coefficient    
1 Introduction

Porous materials (or porous media) are widely used in many fields of applied science and engineering due to their functions such as filtration,separation,sound absorption [1] ,and heat transfer [2] .

Sine porous surfaces can modulate turbulent,the fluid flow through a porous medium is one of the most common topics and has become an interesting separate subject of study [3] . At different porosities and with different Darcy numbers,the effects of porous surface on the modulation of turbulent flow are different. Bruneau and Mortazavi [4] simulated the flow past around the two-dimensional (2D) bluff body which was wrapped by a porous material,and found that the drag coefficient of the bluff body dropped. Naito and Fukagata [5] reported the same trend in conjunction with drag reduction by a simulation on the flow around the threedimensional (3D) cylinder wrapped by a porous medium. The results indicated that the velocity and pressure fluctuations around the cylinder were weakened distinctly as a result of the porous medium. However,Jim´enez et al. [6] found that the turbulent level and the skin friction were to increase rather than to decrease when the turbulent shear flow was over a porous wall. By now,both points of view are complementary,and the detailed mechanism still remains unclear.

To simulate the flow efficiently in porous and fluid media,it is necessary to solve the Darcy equations and the Navier-Stokes equations in these two media simultaneously. This is quite hard to handle since it needs to couple two simulations and to put forward the right condition at the interface of the two media. Jim´enez et al. [6] treated the wall-normal velocity as a transpiration velocity which was proportional to the local wall pressure fluctuations. However,this method is a simplified mathematical method,and cannot recover exactly the impact from the porous wall. Hahn et al. [7] presented a boundary condition at the interface between a permeable block and the turbulent channel flow,and investigated the characteristics of the wall-bounded turbulence with permeable walls. The results showed that the viscous sublayer thickness decreased and the near-wall vortex structures were significantly weakened by the porous wall. In addition, the porous wall reduced the turbulence intensities,the Reynolds shear stress,the pressure and vorticity fluctuations,and the skin-friction. However,their method dealt with the porous medium and the fluid,respectively,and the condition at the interface of the two media is really a troublesome problem.

In fact,each medium can be considered as a porous medium. It is understandable to treat fluid as a porous medium of infinite permeability and solid as a porous medium of zero permeability. In this work,the lattice Boltzmann method (LBM) is used to solve the porous wall-bounded turbulent flow.

The LBM appears to be an attractive approach to compute complex flows [8] ,such as turbulence flow [9, 10] and microscale flow [11] . It has been successfully used in many areas of computational physics,e.g.,aeroacoustics [12] . All in all,the LBM is a rather efficient,robust method which is very easy to implement. Traditional computational fluid dynamics (CFD) methods use to solve the conservation equations of macroscopic properties (mass,momentum,and energy) numerically. The LBM is based on the discrete-velocity Boltzmann equation. In general, propagation and collision of dummy particles are the main interpretation processes in the above procedure.

In this paper,the turbulence modulation in a porous wall-bounded turbulent channel flow is further explored. The physical model and the mathematical formulation are described in Section 2. For the flow inside the porous medium,the volume-averaged method [13] is adopted. We also present how the lattice Boltzmann equation containing the Darcy-Brinkman-Forchheimer acting force term [14, 15] is deduced in this section. Some statistical quantities of the impermeable wallbounded turbulent flow and validations are given in Section 3. The turbulence statistics are presented in Section 4. In Section 5,the turbulent structures of the velocity and vorticity instantaneous field are shown to explain the mechanism for the drag reduction of the porous medium. Lastly,a summary is followed in Section 6. 2 Problem and method 2.1 Governing equation

We consider the turbulent flow between two parallel planes (see Fig. 1). The widths of the computational domain are 2πH ,πH ,and 2H in the streamwise,spanwise,and wall-normal directions,respectively. Here,we set the porous layer close to the two walls. The porous layer’s thickness is δ,which is far less than the distance between the two plates. The porosity of the porous layer is ε,and the permeability is K.

Fig. 1 Porous wall-bounded turbulence model

For the isothermal incompressible flow in the porous medium,according to the BrinkmanForchheimer equation [16] ,the governing equations are given as follows:

where F is the total force which contains the medium resistance and the driving force terms [16, 17], i.e., in which v is the kinematic viscosity,and

For the flow inside the porous medium,this model is averaged over a small spatial volume with dimensions large enough to smooth in the homogeneities at pore scales or small enough to remain the flow dynamics. According to the Ergun empirical formulas [18] ,the structure function Fε and the permeability K are expressed as follows: where d p is the particle diameter of the porous medium.

It is worth pointing out that the porosity tends to be 1.0 and the permeability is approximately infinite except in the porous medium domain. Meanwhile,the force F can be treated as G,and Eq. (1) is equivalent to the standard Navier-Stokes equations in the fluid domain. In other words,the governing equations are the Navier-Stokes equations in general. 2.2 Lattice Boltzmann force model

The lattice Bhatnagar-Gross-Krook (LBGK) [19] is a single relaxation model,which is one of the most widely used lattice Boltzmann equation (LBE) models. In fact,it has the secondorder accuracy in time and space if the flow is incompressible and the numerical viscosity can be absorbed into the physical viscosity [20] .

According to the LBGK model including the forcing term [21] ,this paper deduces the D3Q19 LBE model which solves the 3D porous wall-bounded turbulent flow as follows:

in which the equilibrium distribution function can be expressed as follows: and the forcing term is

The macroscopic density and the velocity are defined by

where c 0 and c 1 are parameters,and v is a temporary velocity expressed by

2.3 Chapman-Enskog expansion derivation

It needs to verify the equivalence between the LBE model and the corresponding NavierStokes equations. With the definition of the equilibrium distribution function,one can easily obtain the following moments:

From the LBE model (see Eq. (4)),the arising macro-dynamical behavior can be found from a multi-scale analysis by use of an expansion parameter λ,which is proportional to the ratio of the lattice spacing to a characteristic macroscopic length. To do so,the following expansions are introduced:

Expanding f i (x + c i δ t,t + δ t ) and applying the above multi-scale expansions to the resulting continuous equation,one can obtain the following equations in the consecutive order of the parameter λ:

where

According to the definitions

and the moment equations (8) and (9),we have

Taking the moments of Eq. (15),we can obtain the following macroscopic equations on the time scale t1 = εt and the space scale x 1 = εx:

where Π (0) is the zeroth-order momentum flux tensor given by

The first-order momentum flux can be simplified by Eq. (18). After some standard algebra,we obtain

where the terms of order O(M 3 ) are neglected. The macroscopic equations on the time scale t2 = λ 2 t are derived by taking the moments of Eq. (16). With the aid of Eq. (18),the final equations can be written as follows: where

Combining the macroscopic equations on t1 and t2 ,one can finally obtain the macroscopic equation as Eq. (1). So far,we have finished the verification by the series expansion. 3 Model validation

The Boltzmann method is easy to be programmed,and has high performance on parallel computing. It is really an efficient way since it is convenient to deal with complex boundaries. Therefore,we simulate the porous wall-bounded turbulent flow by the LBM.

Firstly,to validate the reliability of the current computational procedure,we calculate an impermeable wall-bounded turbulent flow and compare the obtained results with some typical direct numerical simulation (DNS) results. The periodic boundary condition is adopted in the streamwise and spanwise directions,and the correctional bounce-back scheme boundary condition (the 2nd-order accuracy) is used in the wall-normal direction. To save the computational resource and evolve the turbulent flow field as soon as possible,we choose the time development mode and set the initial velocity field [22] as follows:

where the symbol “+” stands for the velocity normalized by the wall shear velocity. Obviously, the velocity field appears as linear-logarithmic distribution. Therefore,the von constant k is 0.4,and the other two parameters B and y w are 5.2,and 11.6,respectively. To arouse the turbulence,we also add perturbation to the initial velocity field. At the first time-step, several vortices are superposed along the streamwise or the spanwise. The random perturbation is also added to the initial velocity field satisfying the divergence-free condition [23] . The perturbation immediately propagates into all the three velocity components,but approximate solenoidality is still maintained [22] .

Table 1 shows the statistical parameters of the wall-bounded turbulence. All the parameters are obtained after the flow field full being developed. InTable 1 ,ymin + is the mesh size in the near wall region normalized by the wall unit,u t ,u c ,and um are the wall shear velocity,the mean centerline velocity,and the mean bulk velocity in the streamwise,respectively,Reτ = u τ H/v is the wall Reynolds number,Rec = Hc H/v is the centerline Reynolds number,Rem = UmH/v is the mean bulk Reynolds number,c f = (2T w )/(ρUm 2 ) is the drag coefficient.

Table 1 Statistical parameters of wall-bounded turbulence

The validation case is carried out in the uniform meshes and the number of the total meshes is about 10% less than that of Kim et al. [24] . The statistics’ errors between the two cases are still within 1%. Therefore,the current method can do simulate the wall-bounded turbulence under lower simulation cost. Through analyzing the velocity statistics (see Fig. 2),in the linear and logarithmic regions,the velocity curve almost coincides with that of Kim et al. [24] .

Fig. 2 Mean velocity in wall coordinate

For the second-order statistics,the root-mean-square velocity fluctuations in the spanwise and wall-normal directions all display good consistency with the results of Kim et al. [24] and Moser et al. [25] ,and the streamwise velocity fluctuation shows a little offset probably on account of two kinds of boundary conditions (see Fig. 3). In addition,the Reynolds stress agrees well with that of Kim et al. [24] (see Fig. 4). In Fig. 4,the total stress is the sum of and .

Fig. 3 Root-mean-square of velocity fluctuation along three directions in wall region

Fig. 4 Shear stress and Reynolds stress profiles

The common high-order statistics data are the skewness and the flatness factors [24] . They all reflect the degrees of the turbulence deviating from the normal distribution. If the Say skewness factor S(ui ) is positive,the positive probability of ui is more likely. The flatness factor is used to the measure frequency of the velocity fluctuations,and it usually characterizes the intensity of the turbulent intermittency. Figure 5 shows the statistical results of the velocity fluctuations. From the figures,we can find that the skewness and flatness factors of the wall-normal velocity coincide well with that of Kim et al. [24] .

In view of the above analysis,it can be confirmed that our calculation based on the LBM is high efficient and reliable for the prediction of statistical quantities of the turbulent flow with a porous surface.

Fig. 5 Skewness and flatness factors in global coordinate
4 Parametric study

The second part of this paper will give a description of the porous wall-bounded turbulence. There are mainly two parameters for the porous walls,i.e.,the permeability K and the porosity ε. Here,we replace the permeability with the Darcy number expressed by Da = K/H 2 . In addition,the thickness δ of the porous walls is rather small compared with the distance between both walls,and we assume that

4.1 Effects of Darcy number

Firstly,we discuss the turbulent transport at different Darcy numbers (the porosity of wall is 0.2). We set several cases listed inTable 2 ,in which Case 0 stands for the conventional wall-bounded turbulence and Cases 1,2,and 3 stand for the porous wall-bounded turbulence at different Darcy numbers. The statistical results obtained from Cases 1 and 2 show that the mean velocities and the central velocities in the streamwise direction are all lower than those obtained form Case 0 while higher than those obtained from Case 3.

Table 2 Statistical parameters at different Darcy numbers

Figure 6 shows the streamwise mean velocities along the normal direction. We can obviously find that the mean velocity increases in the central region and decreases near the wall when the Darcy number decreases. We can find that there exists a cross point on the curves of the mean profile. In fact,it is significant in the flow of the turbulent boundary layer. Usually,its location stands for the interface area between the viscous sublayer and the logarithmic law layer.

Fig. 6 Mean velocity in stream direction

The porous walls can modulate the velocity fluctuations. Figures 7(a),7(b),and 7(c) show the distributions of the streamwise,wall-normal,and spanwise root-mean-square velocity fluctuations normalized by the wall shear velocity at different Darcy numbers. When the Darcy number increases,the streamwise velocity fluctuation decreases in the near wall region and increases in the far wall area; and the spanwise and wall-normal velocity fluctuations in all region increase obviously. In addition,the Reynolds stress shows the same variation trend (see Fig. 7(d)).

Fig. 7 Root-mean-square velocity fluctuations and Reynolds shear stress

Figure 8 shows the skewness and flatness factors of the velocity fluctuations at different Darcy numbers. From the figure,we can find that the flatness factors of the streamwise and wallnormal velocity fluctuations decreases when the Darcy numbers increases (see Figs. 8(a),8(b), and 5),and the skewness factor doesn’t display a change obviously. Under the same situations, the skewness factor of the streamwise velocity fluctuation decreases,while the skewness factor of the wall-normal velocity fluctuation increases. Therefore,we can draw a conclusion that the porous walls can actually weaken the sweeping movement in the near wall region.

Fig. 8 Skewness and flatness factors

Some high-order statistics of the Reynolds stress u v are also analyzed to investigate the influence of the porous wall on the Reynolds stress. From Figs. 8(c) and 8(d),when the Darcy number becomes larger,the amplitude of the skewness factor in the near wall region gets smaller. Moreover,the differences among the flatness factors of the Reynolds stress from different porous walls are obvious. This indicates that the Darcy number has a remarkable impact on the Reynolds stress in the whole flow field. 4.2 Effects of porosity

According to Section 2.1,we know that the force term F in the porous region can be expressed by the porosity,the Darcy number,etc. To study how the porosity affects the wallbounded turbulence,we firstly discuss the functional relation between the fluid drag and the porosity of the porous walls. Substituting Eq. (3) into Eq. (2),we can get

Obviously,there exists a nonlinear functional relation between F and ε. According to the cases of this work,we can assign the parameters as follows (all the parameters are set in the lattice unit):

Now,plot three graphs (Fx ~ ε) at different Darcy numbers (see Fig. 9). As shown in Fig. 9, when the Darcy number is 5×10 −2 ,the force in the streamwise direction increases when the porosity increases; when the Darcy number is 5×10 −4 (hereinafter it is referred to as the criticalDarcy number),the force increases at first,and decreases later; and when the Darcy number is 5×10 −6 ,the force is monotonically decreasing.

Fig. 9 Functional relations between Fx and ε at different porosities

Now,we design some new cases according to the above functional relation between the drag and the porosity (see Table 3 ). Cases 4 and 5 are under different porosities,and their Darcy numbers are both 5×10 −4 . It is easy to find that Case 4 displays a better drag reduction effect than Case 5. By comparing Case 3 with Case 6,we can find that the porous walls at larger porosity get better drag reduction effects. According to these simulations,the maximum drag reduction coefficient is 4.5% under Case 6.

Table 3 Statistical parameters at different porosities

Figure 10 shows the streamwise mean velocities for four Cases 3,4,5,and 6. Case 4 reveals the largest mean velocity in the near wall region,and Case 6 displays the largest velocity in the far wall area. It is worth mentioning that with the same Darcy number,increasing porosity doesn’t mean obtaining larger mean velocity.

Fig. 10 Mean velocity in stream direction

We examine the root-mean-square velocity fluctuations and Reynolds stresses for different porous wall-bounded turbulent flows. Figure 11(a) shows that Case 4 has smaller velocity fluctuations than Case 5 in all directions,and has smaller Reynolds stress than Case 5 (see Fig. 11(b)).

Fig. 11 Root-mean-square velocity fluctuations ((a) and (c)) and Reynolds shear stresses ((b) and (d)) normalized by wall shear velocity

From Case 2,Case 4,and Case 5,we can find that the drag coefficient increases firstly,and then decreases when the porosity increases with the same Darcy number. This phenomenon can be explained by the root-mean-square velocity fluctuations and Reynolds stresses. When the Darcy number is small enough,e.g.,Da 5 × 10 −6 ,the streamwise velocity fluctuation decreases and the spanwise and wall-normal velocity fluctuations increase when the porosity increases (see Fig. 11(c)). Even so,the Reynolds stress in the fluid region is reduced (see Fig. 11(d)). 5 Analysis of velocity and vorticity instantaneous fields

Figure 12 shows the contours of the streamwise instantaneous velocity in the xz-plane for Cases 0,1,and 6,respectively. By comparing Fig. 12(b) with Fig. 12(a),the contour of Fig. 12(b) displays more smaller structures,which appears as distortions of the large-eddy contours. However,the streamwise coherence of Case 6 in Fig. 12(c) displays stronger velocity correlation in the streamwise direction. The low-speed streaks of the near-wall flow structures in Fig. 12(c) are considerably increased,and they are longer and more regular in the geometric profile than those in Fig. 12(a). Consequently,this is a typical feature of drag-reduced flows,which is independent of the actual cause of drag reduction. On the contrary,Fig. 12(b) does not reveal these features. The streamwise instantaneous velocity contours in the xz-plane in the far wall area are shown in Fig. 13. Figure 13(c) shows more high-speed streaks than Fig. 13(b). That is to say,Case 6 has a larger mean velocity in the core area than Case 1,which is an intuitive behavior to drag reduction.

Fig. 12 Contours of velocity in stream direction at xz-plane (y + =12)

Fig. 13 Contours of velocity in stream direction at xz-plane (y + =60)

We have already known that shear is the major characteristic in the turbulence boundary layer. Since the conventional vorticity criterion (ω = rot V ) is disabled in this kind flow,we distinguish the vortex structures in the near wall region by the Q-criterion [26] . The criterion is defined as

where Wij and Sij are,respectively,the symmetric and skew-symmetric velocity gradient tensors expressed by

Figure 14 shows the vortex structures in the near wall region of conventional and the porous wall-bounded turbulence. From Case 6,we can find that the coherent structures,the scales, and the attack angles obtained from Case 6 are all reduced with respect to those obtained from Case 0. On the contrary,Case 1 exactly gets enhancive coherent structures. According to Kolmogorov’s K41 theory [27] ,the dissipative eddies (small scale magnitude) always dissipate the energy from large scale vortexes. Obviously,since the turbulence of Case 6 has more micro vortexes than those of Case 1,the former will dissipative less energy,and then the drag reduction can be achieved.

Fig. 14 Vortex structure in near wall region via Q-criterion (Q = 0.006)
6 Conclusions

We investigate the characteristics of turbulent flow and the turbulence modulations of wallbounded turbulence in the presence of porous walls by means of a direct numerical simulation by the LBM. At different porosities and with different Darcy numbers,pronounced turbulence modulations are observed. From the analysis,the following conclusions can be made:

(i) The porous surface can modulate the turbulence intensity and the Reynolds shear stresses in near wall region,which can affect the streamwise coherence of the near wall flow structures. In the view of physics,the porous medium is essentially to generate an intermediate flow,which controls the vortex flows and weakens the boundary layer.

(ii) The mean velocity increases when the Darcy number decreases. In addition,when the Darcy number decreases,the streamwise velocity fluctuation increases in the near wall region, and the growth trend slows down. As a result,the turbulent anisotropy weakens the wall-normal and spanwise velocity fluctuations and the Reynolds stresses.

(iii) When the Darcy number is in the order of O(10 −4 ),the streamwise mean velocities and velocity fluctuations are indeterminate. The present results show that very small porosity may also result in weak velocity fluctuations and Reynolds stresses in the near wall region.

(iv) The drag coefficient is damped in the porous wall-bounded turbulent flow when the Darcy number is in the order of O(10 −6 ). The streamwise low-speed streaks become longer and more regular than those of the impermeable wall-bounded flow,and appear stronger velocity correlation in the streamwise direction. At the same time,the coherent structures of the vorticity field and the scale and attack angles of vortex in the near wall region are reduced.

References
[1] Oliva, D. and Hongisto, V. Sound absorption of porous materials accuracy of prediction meth- ods. Applied Acoustics, 74, 1473-1479 (2013)
[2] Ayuttaya, S. S. N., Chaktranond, C., and Rattanadecho, P. Numerical analysis of electric force influence on heat transfer in a channel flow (theory based on saturated porous medium approach). International Journal of Heat and Mass Transfer, 64, 361-374 (2013)
[3] Philip, J. R. Flow in porous media. Annual Review of Fluid Mechanics, 2, 177-204 (1970)
[4] Bruneau, C. H. and Mortazavi, I. Passive control of the flow around a square cylinder using porous media. International Journal for Numerical Methods in Fluids, 46, 415-433 (2004)
[5] Naito, H. and Fukagata, K. Numerical simulation of flow around a circular cylinder having porous surface. Physics of Fluids, 24, 117102 (2012)
[6] Jiménez, J., Uhlmann, M., Pinelli, A., and Kawahara, G. Turbulent shear flow over active and passive porous surfaces. Journal of Fluid Mechanics, 442, 89-117 (2001)
[7] Hahn, S., Je, J., and Choi, H. Direct numerical simulation of turbulent channel flow with perme- able walls. Journal of Fluid Mechanics, 450, 259-286 (2002)
[8] Aidun, C. K. and Clausen, J. R. Lattice-Boltzmann method for complex flows. Annual Review of Fluid Mechanics, 42, 439-472 (2010)
[9] Dong, Y. H. and Sagaut, P. A study of time correlations in lattice Boltzmann-based large-eddy simulation of isotropic turbulence. Physics of Fluids, 20, 035105 (2008)
[10] Dong, Y. H., Sagaut, P., and Marie, S. Inertial consistent subgrid model for large-eddy simulation based on the lattice Boltzmann method. Physics of Fluids, 20, 035104 (2008)
[11] Spaid, M. A. A. and Phelan, F. R. Lattice Boltzmann methods for modeling microscale flow in fibrous porous media. Physics of Fluids, 9, 2468-2474 (1997)
[12] Deng, Y. Q., Tang, Z., and Dong, Y. H. Lattice Boltzmann method for simulation propagating acoustic waves. Chinese Journal of Computational Physics, 30, 808-814 (2013)
[13] Breugem, W., Boersma, B., and Uittenbogaard, R. The influence of wall permeability on turbulent channel flow. Journal of Fluid Mechanics, 562, 35-72 (2006)
[14] Luo, L. S. Unified theory of lattice Boltzmann models for nonideal gases. Physical Review Letters, 81, 1618-1621 (1998)
[15] Martys, N. S. Improved approximation of the Brinkman equation using a lattice Boltzmann method. Physics of Fluids, 13, 1807-1810 (2001)
[16] Nithiarasu, P., Seetharamu, K., and Sundararajan, T. Natural convective heat transfer in a fluid saturated variable porosity medium. International Journal of Heat and Mass Transfer, 40, 3955-3967 (1997)
[17] Seta, T., Takegoshi, E., and Okui, K. Lattice Boltzmann simulation of natural convection in porous media. Mathematics and Computers in Simulation, 72, 195-200 (2006)
[18] Ergun, S. Fluid flow through packed columns. Chemical Engineering Progress, 48, 89-94 (1952)
[19] Shi, B. C., Guo, Z. L., and Wang, N. C. Lattice Bhatnagar-Gross-Krook simulations of turbulent natural convection in a cavity. Chinese Physics Letters, 19, 515-517 (2002)
[20] Qian, Y. H., D'Humières, D., and Lallemand, P. Lattice BGK models for Navier-Stokes equation. Europhysics Letters, 17, 479-484 (1992)
[21] Guo, Z. L., Zheng, C. G., and Shi, B. C. Discrete lattice effects on the forcing term in the lattice Boltzmann method. Physical Review E, 65, 046308 (2002)
[22] Lammers, P., Beronov, K. N., Volkert, R., Brenner, G., and Durst, F. Lattice BGK direct numer- ical simulation of fully developed turbulence in incompressible plane channel flow. Computers & Fluids, 35, 1137-1153 (2006)
[23] Tuckerman, L. S. Divergence-free velocity fields in nonperiodic geometries. Journal of Computational Physics, 80, 403-441 (1989)
[24] Kim, J., Moin, P., and Moser, R. Turbulence statistics in fully developed channel flow at low Reynolds number. Journal of Fluid Mechanics, 177, 133-166 (1987)
[25] Moser, R. D., Kim, J., and Mansour, N. N. Direct numerical simulation of turbulent channel flow up to ReT= 590. Physics of Fluids, 11, 943-945 (1999)
[26] Hunt, J. C. R., Way, A., and Moin, P. Eddies, stream, and convergence zones in turbulent flows. Proceedings of the 1988 Summer Program, NASA, California, 193-208 (1988)
[27] Frisch, U. Turbulence, Cambridge University Press, Cambridge, 100-106 (1996)