You are on page 1of 51

Accepted Manuscript

Multiple-relaxation-time lattice Boltzmann simulation of non-Newtonian flows past a rotating circular cylinder Keivan Fallah, Morteza Khayat, Mohammad Hossein Borghei, Atena Ghaderi, Ehsan Fattahi PII: DOI: Reference: To appear in: S0377-0257(12)00067-5 10.1016/j.jnnfm.2012.03.014 JNNFM 3331 Journal of Non-Newtonian Fluid Mechanics

Please cite this article as: K. Fallah, M. Khayat, M.H. Borghei, A. Ghaderi, E. Fattahi, Multiple-relaxation-time lattice Boltzmann simulation of non-Newtonian flows past a rotating circular cylinder, Journal of Non-Newtonian Fluid Mechanics (2012), doi: 10.1016/j.jnnfm.2012.03.014

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Title Multiple-relaxation-time lattice Boltzmann simulation of non-Newtonian flows past a rotating circular cylinder

Keivan Fallah 1, Morteza Khayat2, Mohammad Hossein Borghei3, Atena Ghaderi4, Ehsan Fattahi5

Department of Mechanical Engineering, Sari branch, Islamic Azad University, Iran; Keyvan.Fallah@gmail.com (Tel: +989113279370, Fax: +98 (21) 77896626)

Department of Mechanical Engineering, Science and research branch, Islamic Azad University ,Tehran, Iran; Mkhayat@srbiau.ac.ir

Faculty of Mechanical Engineering, Iran University of Science and Technology, Tehran, Iran; Borghei@iust.ac.ir

Department of Mechanical Engineering, Ayatollah Amoli branch, Islamic Azad University, Amol, Iran; A.ghaderi64@gmail.com
4

Faculty of Mechanical Engineering, Babol University of Technology, Babol, Iran; E.Fattahi@stu.nit.ac.ir

Abstract The flow field around a rotating circular cylinder is studied numerically using Lattice Boltzmann method via multi-relaxation-time approach. Simulations are performed at a fixed Reynolds number of 100 while dimensionless rotational ratio () and power-law index (n) range as, 02.5 and 0.4n1.8, respectively.The effects of dimensionless rotational ratio and the Power-Law indexs on the flow field, mean drag and lift coefficients, Strouhal number and pressure coefficient are investigated in detail. To verify the simulation, the results are compared to previous experimental and numerical data.

Keywords lattice Boltzmann method; Multiple-relaxation-time; Power-law fluids; Drag coefficient; Lift coefficient; Rotating circular cylinder.

1. Introduction Flow pattern and its characteristics around an immersed bluff body have been investigated by many researchers because of the complexities and practical importance of these flows. The principal issues of this classical flow configuration are wake patterns, boundary-layer separation, vortex shedding, and external forces acting on the body (drag and lift characteristics).The results show that the flow over a cylinder depends on a large number of parameters such as the nature of the far flow field (uniform, shear or extensional), type of fluid (compressible or incompressible, Newtonian or non-Newtonian), unconfined or confined cylinder, stationary or rotating cylinder, etc. Newtonian fluid flows over circular cylinders arise in many engineering processes, such as in transmission lines, heat exchangers, offshore structures and pipeline, marine cables, etc. Lots of detailed information exists on the Newtonian flow past a circular cylinder [1-6]. Non-Newtonian flow over a circular cylinder has been investigated less extensively than Newtonian flow. Nevertheless, it can be found in several engineering applications, for example, filtration devices, production of fiber-reinforced composites, catalytic chemical reactors, lungs modeling, membrane based separation modules and so on. Therefore, more systematic study on the non-Newtonian flow is needed. Coelho and Pinho [7-9] conducted experimental studies to investigate the non-Newtonian flow pattern around circular cylinder. They reported considerable results in the three part series on the identification of several flow regimes and the related values of the transition Reynolds number, and on the pressure measurements in well-characterized Newtonian, shearthinning fluids with weak and strong viscoelasticity. Sivakumar et al. [10] investigated numerically the effect of power-law index on the critical Reynolds numbers which denotes the onset of wake instability, Strouhal number and drag coefficient over a wide range of
3

power-law index (0.3n1.8).Bharti et al. [11] studied the steady two-dimensional fluids over a circular cylinder using a semi-implicit finite volume method. They investigated the roles of Reynolds number and power law index on the flow patterns in the range of 5 Re 40 and the power-law index 0.6 n 2. Patil et al. [12] examinedanumericalresearch about the steady flow of power-law fluids past two circular cylinders in a tandem arrangement. They investigated the influences of the power-law index (0.4 n 1.8), Reynolds number (1 Re 40), and the gap ratio between the two cylinders (2 G 10) on flow variables such as drag coefficient, pressure coefficient and streamline, etc., in detail.Sivakumar et al. [13] performed a study onthe power-law fluid flow past an unconfined elliptic cylinder. Numerical simulations are performed for different Reynolds numbers (0.01 Re 40), in the range of aspect ratio of the elliptic cylinder 0.2 E 5, and the power-law index 0.2 n 1.8. Patnana et al [14] carried out a numerical study of unsteady flow of power-law fluids over a circular cylinder at low Reynolds numbers while Reynolds number and power-law index varied in the extent of 40 Re 140 and 0.4 n 1.8, respectively. Moreover, Saroj et al. [15] conducted a study on uniform flow past a rotating circular cylinder. Their simulations were carried out over the range of 0.1Re 40, power-law index between 0.2 and 1(0.2n1), and dimensionless rotational ratio () was in the range of 0 6. Currently, a wide range of computer-intensive problems in fluid mechanics can be solved quite efficiently using numerical methods such as Finite Difference Method (FDM), Finite Volume Method (FVM), Finite Element Method (FEM), Boundary Element Method (BEM), Spectral Method (SM), and Lattice Boltzmann Method (LBM), with the latter method being the most recent one in this long list [16]. The Lattice Boltzmann Method (LBM) [17-19] is a numerical method for solving the incompressible Navier-Stokes equations. LBM has been

successfully extended to various types of flows, such as natural convection [20], nanofluid [21,22], porous media [23-26], multiphase flow [27], unsteady flow [28-30], etc. In our recent studies, Lattice Boltzmann method is applied to investigate two dimensional laminar shear flow past a rotating circular cylinder [29] and two stationary equal-sized circular cylinders in tandem arrangement [30].Quantitative information about the flow variables such as drag and lift coefficients, pressure coefficient and vorticity distributions on the cylinders was highlighted. Simulations were performed at a fixed blockage ratio (B = 0.1) while the Reynolds number, nondimensional shear rate (K) and absolute rotational speed (ratio of the peripheral velocity to the inflow velocity) range as, 80 Re 180, 0 K 0.2 and -2 2, respectively [29]. Similarly, analogous studies on have been reported for Reynolds number (Re=200) while the nondimensional shear rate (K) and gap spacing range as 0 K 0.2 and 0.5 g* 6 [30]. In recent years, some efforts have been made to extend the LBM for simulation of nonNewtonian fluid flows [3140]. This is because of the ability of LBM for tackling problems in non-Newtonian fluid mechanics, where the constitutive behavior of the fluid is often found to be nonlinear. One can mention, for example, the Lattice Boltzmann method in which the rate-of-deformation tensor can be computed locally thereby eliminating the need for computing the troublesome space derivatives of the velocity field [41]. The other advantages of the LBM are due to the fact that the solution for the particle distribution functions is explicit, easy for parallel computation, and simple implementation of boundary conditions on curved boundaries. The simplest Lattice Boltzmann equation (LBE) is the lattice Bhatnagar-Gross-Krook (BGK) equation based on a single-relaxation time (SRT) approximation [42]. Due to its extreme simplicity, the BGK model is the most popular one. Although, this simplicity comes at the expense of numerical instability [43] and inaccuracy in implementing boundary conditions
5

[44]. To overcome these deficiencies in the BGK models, the multiple-relaxation time (MRT) model was introduced by dHumieres [45]. For non-Newtonian flows, the viscosity is not just a function of fluid properties and it depends on the local strain-rate tensor. Differentiation of the velocity field reduces the numerical accuracy and can cause instability in the solution process. As mentioned earlier, one of the advantages of the LBE is that the viscous stresses can be calculated locally. Chai et al. [46] presented a multiple-relaxation-time lattice Boltzmann model for non-Newtonian fluid flows, which can be considered as an extension to some available works. In computational fluid dynamics, the ability to handle complex geometries accurately and efficiently has been the primary discussion. Various methodologies have been proposed for simulating the flow over complex geometries while using LBM. For example, Filippova and Hanel [47] proposed a second-order accurate boundary condition for curved geometry which was developed in conjunction with the use of Cartesian grids in order to retain the advantages of the LBE method. Mei et al. [48] solved LBE in curvilinear coordinates using a finite difference method and improved its numerical stability. Lallemand and Luo [49] composed interpolation scheme and the bounce-back scheme to treat a moving curved boundary by the Lattice Boltzmann Method. In this study, the Lattice Boltzmann Method with multiple-relaxation-time (MRT) collision model is applied to simulate 2-D non-Newtonian flow passing a rotating circular cylinder. The study was carried out at a various range of power-law index (0.4n1.8) and dimensionless rotational ratio (02.5) at Reynolds number of 100. The present study is an attempt to investigate the effects of power-law index and rotational speed

2. Numerical methods
6

2.1. Multiple-relaxation-time lattice Boltzmann equation (MRT-LBE) In this paper, the lattice Boltzmann equation using the MRT collision model with nine velocities (D2Q9 model) is considered [43]. The evolution equation of the model is as:
f ( x + e k t , t + t ) f ( x, t ) = M -1 S M (f ( x, t ) f eq ( x, t ) )

(1)

where f =(f 0 ; f1 ; f 2 ; f 3 ; f 4 ; f 5 ; f 6 ; f 7 ; f8 ) T and T denotes the transpose operator. f (x, t ) is 9dimensional vectors for the distribution functions, f eq (x, t ) is the equilibrium distribution function of f (x, t ) , and computed as:

f keq = w k (1 +

3 9 3 e .u + 4 (e k .u) 2 2 u.u ) 2 k 2c 2c c

(2)

where c=x/t is the lattice speed, x and t are the lattice cell width and time step size, respectively. Here, t is chosen to be equal to x, thus c=1. The macroscopic fluid variables, density, velocity, and pressure are calculated as follows:

= fk
k =0

(3)

u=

k =0

fk

(4)

p = Cs2
where C s = c / 3 is the speed of sound. As shown in figure 1, ek denotes the particle velocity and is defined as:

(5)

(0, 0) k =0 k = 1, 2, 3, 4 e k = c (cos , sin ) = ( k 1) / 2 2c (cos , sin ) = ( k 5) / 2 + / 4 k = 5, 6, 7,8


Also, wk is a weighting factor in the kth direction in the discrete velocity as follows:
4 / 9 w k = 1 / 9 1 / 36
k =0 k = 1, 2, 3, 4 k = 5, 6, 7,8

(6)

(7)

The evolution process of the MRT model also consists of two parts, i.e., collision and streaming. Collision and streaming steps are as follows, respectively:

f k (x, t ) = f k (x, t ) + M1 S [R(x, t ) Req (x, t )]


f k (x+e k t , t + t ) = f k (x, t )

(8)

(9)

where f k (x, t ) denotes the density distribution function after collision.


M is a 99 matrix which transforms f k and f keq into the moment space with R = M f and

R eq = M f eq where R and Req are vectors of macroscopic variables.


The distribution functions in moment space are given by:
e jx R = q x jy q y p xx p xy 1 1 4 1 4 2 0 1 = 0 2 0 0 0 0 0 1 0 0 1 1 1 1 1 1 1 f 0 2 f 1 1 f 2 1 f 3 1 f 4 = Mf 1 f 5 1 f 6 0 f 7 1 f 8

1 1 1 2 2 2 2 2 2 1 1 1 0 1 0 1 1 1 0 1 2 1 0 2 0 0 1 0 0 1 2 1 0 1 1 1 1 1 1 1 1 1 0 0 0 1 1 1

(10)

where, is the fluid density, is related to the square of the energy e, jx=u and jy=v are the mass flux in two directions, qx and qy correspond to the energy flux in two directions, and pxx and pxy correspond to the diagonal and off-diagonal component of the viscous stress tensor. The equilibrium distribution functions in moment space are defined as:
eq eq eq eq Req = ( 0 , e eq , eq , j x , q x , j y , q y , p xx , p xy )T

(11)

where the equilibrium values in the above equation can be written as:

e eq = 2 + 3 ( j x2 + j y2 )
eq qx = j x

eq = 3 ( j x2 + j y2 ) 0
(12)

q eq = j y y
eq p xx = ( j x2 j y2 ) 0 eq p xy = j x j y 0

The constant 0 is the mean density in the domain and is set to be unity. Sj are the relaxation rates that are the diagonal elements of the matrix S .

S = diag (0, S1 , S2 ,0, S4 ,0, S6 , S7 , S8 )

(13)

Due to mass and momentum conservation before and after collision, collision rates S0, S3 and S5 are set zero. In the SRT collision model, the collision operator S is a diagonal matrix whose non-zero elements are 1 , where is the dimensionless relaxation rate. But in the MRT model, the collision operator is a diagonal matrix whose elements values S1, S2, S4 and S6 can be chosen more flexibly in the range 0<Sj<2. Lallemand and Lou [43] recommended that for stability reasons the values should be chosen somewhat more than unity. In this study the relaxation rates S1=S2=S4 =S6=1.1 are used. For giving a consistent dynamics viscosity, relaxation rates S7 and S8 have to be equal S7= S8. S7 and S8 are related to the dynamic viscosity by [43]:

= C s2 t (

1 1 1 1 ) = C s2 t ( ) S7 2 S8 2

(14)

Since strain rate tensor and viscosity for non-Newtonian fluids vary with position in a flow domain, it is needed to determine the shear rate at each point in the simulation. This can be determined by the strain rate tensor:

= ( u + u )

1 2

(15)

One of the advantages of the LBE is that the viscous stresses can be calculated locally. The strain rate tensor in MRT model can be derived with the following equation [33]:
8 8 1 ei ei (M -1 S M )ij [ f j ( x, t ) f jeq ( x, t )] 2 2 C S t i = 0 j =0

(16)

2.2. The power-law model

Apparent viscosity of a power-law fluid, , is determined as:

= 0

n 1

(17)

where 0 is the zero-shear viscosity and n is the power-law exponent. The case n< 1 correspond to shear-thinning (pseudoplastic) fluids, whereas n> 1 correspond to shearthickening (dilatant) fluids and n=1 recovers the Newtonian behavior. Strain rate is related to the symmetric strain rate tensor,
DII =

, =1

(18)

by

= DII
where here L = 2 for a two-dimensional simulation. The non-Newtonian MRT model is implemented using the following procedure:
10

(19)

(1) All fk are initialized by setting each of the distribution functions equal to their equilibrium value: fk(x, t = 0) = fkeq(x, t = 0). The equilibrium distribution function fkeq can be determined from Eq. (2), with the initial constant density and zero velocity. (2) The fluid density and velocity are calculated from Eq. (3) and (4). (3) S7 and S8 are determined from equations (16), (18), (19), (17) and (14). (4) The collision step is conducted based on Eq. (8). (5) The curve boundary treatment is carried out (6) The streaming step is performed based on Eq. (9). (7) Inflow, outflow, and sides boundary conditions are operated. (8) Proceed through the next time-step starting at step 2.
2.3. Curved boundary treatment

The lattice nodes on the solid and fluid side are denoted as xb and x f , respectively, in figure 2. The black small circles on the boundary, x w , denote the intersections of the wall with various lattice links. The boundary velocity at x w is u w . The fraction of an intersected link in the fluid region is , which is defined as:
x f xw x f xb

(20)

The standard (half-way) bounce back no-slip boundary condition always assumes a delta value of 0.5 to the boundary wall (figure 3a). Due to the curved boundaries, delta values in the interval of (0, 1] are now possible. Figure 3b shows the bounce back behavior of a surface with a delta value smaller than 0.5 and figure 3c shows the bounce back behavior of a wall with delta bigger than 0.5. In all three cases, the reflected distribution function f k (x, t + t ) at
11

x f is unknown. Since the fluid particles in the LBM are always considered to move one cell
length per time step, the fluid particles would come to rest at an intermediate node xi . In order to calculate the reflected distribution function in node x f , an interpolation scheme has to be applied. For treating velocity field in curved boundaries, the technique is based on the method reported by Guo et al. [50].
2.3.1. Velocity in curved boundary condition

In order to calculate the velocity on a curved boundary, it is needed to establish f k (xb , t + t ) with respect to known information and parameters. Filippova and Hnel [47] essentially proposed a method using the linear interpolation:

f k (xb , t + t ) = (1 ) fk (x f , t + t ) + fk (xb , t + t ) 2
where

3 w (x f , t + t )ek .uw c2 k

(21)

f k (xb , t + t ) = f keq (x f , t + t ) +
and

3 w ( x f , t + t )e k . ubf u f c2 k

(22)

u bf = u ff ,

2 1 1 , if 0 < 2 m 2

(22a)

ubf = (1

3 3 2 1 1 )u f + uw , = , if < 1 1 2 2 2 m + 2

(22b)

where uw , and u bf are the velocity at the wall, weighting factor, the imaginary velocity for interpolations, respectively. It should be noted that ek e k .
12

3. Problem description

Figure 4 shows the flow geometry, coordinate system and computational domain. The freestream with a constant velocity profile, u=U, passes over a circular cylinder that has a diameter of D and rotates with the counterclockwise rotational speed . In order to eliminate the effect of the boundaries, the center of the cylinder is located far enough from the boundaries of the flow field. The center of the cylinder is 8D away from the inlet while it is located 5D far from the upper and lower boundaries. At the inflow, i.e. the left boundary, fluid with constant uniform velocity U in x-direction is injected into the domain. Hence, Zou and He [51] boundary conditions are applied for inlet. On the top and bottom boundaries symmetry boundary condition is applied by Mei et al. [48]. The extrapolation scheme presented in Yu et al. [52] is used for the outlet boundary condition which is placed 17D away from the center of the cylinder, as:

f i ( N x , j ) = 2 f i ( N x 1, j ) f i ( N x 2, j )
where Nx is the number of lattices in the x-direction.

(23)

For all simulations, the Reynolds number, based on the cylinder diameter and average inlet velocity, is kept constant at Re = 100, while power-law index (n) and dimensionless rotational speed () are varied from 0.4 to 1.8 and 0 to 2.5, respectively. In the present study, for calculating forces acting on the circular cylinder, the method suggested by Mei et al. [53] has been applied. It is important to introduce some dimensionless parameters. The Reynolds number often used for power-law fluid flows through circular cylinder is defined as:
Re =
2 U n D n 0

13

where U and D are inflow velocity and cylinder diameter, respectively. Surface pressure coefficient, CP:

CP =

(P P ) 1 2 U 2

where P is the local pressure and P is the free-stream pressure. The drag coefficient, CD, and the lift coefficient, CL.

CD =

1 2 U D 2

Fx

, CL =

Fy

1 2 U D 2

Fy and Fx are force components in y and x directions. The Strouhal number, St:

St =

fs D U

where fs is the vortex shedding frequency from the cylinder. fs is obtained by analysis of the time history of the lift coefficient. Dimensionless time, t*:

t* =

2 U t D

Dimensionless rotational speed, :

D . 2U

14

4. Results and Discussion 4.1. Validation


To validate the present study two different problems are considered. First, uniform flow past a rotating circular cylinder is simulated at Re=200, = 0.5. The evolution of wake flow pattern around the cylinder at Re = 200, for t*= 3, 7, 9 and 11 is represented in figure 5. The figure shows the obtained streamlines of the present study in comparison with the experimental results of Coutanceau and Menard [3] and numerical results of Yan and Zu [4]. The comparison indicates that the results have good agreement with experimental data in terms of the formation of wakes. Second, the result of Newtonian and non-Newtonian flow around a stationary and rotating circular cylinder has been represented. The time-averaged drag coefficient (C D ) , the time-averaged lift coefficient (C L ) and the Strouhal number (St) at different dimensionless rotational ratio () for Newtonian flow are presented in Table 1 and are compared with the numerical results of Yan and Zu [4] at a fixed Re = 200. The results show excellent agreement. A grid independency study has been carried out to determine the best mesh. To do this, different computations are performed in the case of Re = 200, =1.0 and n=1.0 on different grids. Four successively grids as 600 200, 900 300, 1200 400 and 1500 500 are considered. As shown in Table 1, a 1200 400 grid could be selected to optimize the relation between the accuracy and the computation time. It can be seen that the maximum difference between the values of time averaged drag in lift coefficients and Strouhal number obtained for the 1200 400 case and the nest 1500 500 case is less than 1.58, 0.59 and 0.1 %, respectively. Therefore, the grid of 1200 400 nodes is selected for all simulations done in present work.

15

In addition, the result of time-averaged drag coefficient (C D ) and Strouhal number for different power-law index (n) at different Reynolds numbers for stationary cylinder are compared with the other numerical data in Table 2. It can be clearly seen that the present numerical approach is in good agreement with the previous ones. After validating the numerical method with previous studies, the behavior of flow and the effects of Reynolds number, power-law index and dimensionless rotational ratio on flow characteristics are discussed.

4.2. Flow and vorticity pattern


In order to obtain insight into the detailed flow structures, initially, the patterns of vorticity contours at different power-law index and dimensionless rotational ratios for a fixed Reynolds number (Re=100) are presented in Figures 6 and 7. Due to the large quantity of data, some results are omitted. Figure 6 shows the patterns of vorticity contours in the vicinity of the stationary cylinder for Re=100 at different power-law index (0.4n1.8). In the case of no rotation, = 0, vortices of the positive and negative signs are alternately shed and then they move downstream. It can be seen that when the fluids behavior changes from Newtonian (n=1) to shear-thickening (n>1), as n increases, the length and the width of the wakes increase. Unlike these cases for the change in the fluids behavior from Newtonian (n=1) to shear-thinning (n<1), this behavior is reversed (i.e. as n decreases the length and the width of the wakes decrease). This observation is in line with Patnana et al. [14]. Figure 7(a), (b) and (c) illustrate the effect of rotation on the vorticity contours for different speed ratios at n=0.4, n=1.0, and n=1.8, respectively. In all the cases, with increasing dimensionless rotational speed , all the vortices become narrower compared to the case of no rotation, leading to decrease in the lateral width of wake and vortex shedding pattern disappears and eventually attains the steady state. In order to clarify
16

the existence of steady limiting behavior in the flow field, the vortices patterns at bigger dimensionless rotational speeds have been shown in Figure 8. It can be clearly seen that flow still remains steady and no shedding happens at >2.5.

4.3. Drag and lift coefficient


Figures 912 display some selected time evolutions of the drag and lift coefficients and illustrate the influence of the dimensionless rotational speed () and flow behavior index (n). Figure 9 shows the time histories of drag and lift coefficients at different power-law index for =0. With increaseing the value of the power-law index ,the amplitude of periodicity of drag and lift decreases. The change in the fluids behavior from Newtonian to shear-thinning significantly augments the frequency of the fluctuations of drag and lift coefficients; an opposite trend is seen for shear-thickening fluids. Figures 10, 11 and 12 show the effect of dimensionless rotational speed () on time histories of drag and lift coefficients for n=0.4, 1.0 and 1.8, respectively. Graphs report that the period time of the fluctuations decreases as rotational speed increases. Also, for the cases where voretx-shedding disappears and flow becomes steady, the amplitude of periodicity of drag and lift coefficient is obviously zero. For better discussion, the quantitative variation of the mean drag coefficient, the mean lift coefficients and the Strouhal number versus dimensionless rotational speed () at different power-law index (n) are shown in figures 13, 14 and 15, respectively. Figure 13 demonstrates the variation of the mean drag coefficient with dimensionless rotational speed () at different power-law index (0.4n1.8). The diagram shows that the mean drag coefficient decreases with the increasing dimensionless rotational speed at the same power-law index (n). The maximum decrease in the drag values is seen for shear-thickening (n>1), especially for n=1.8. Generally, the total
17

drag forces are composed of both the pressure and friction forces. As increases, the pressure drag decreases with the increase in the friction drag, resulting in the net decrease in the total drag force. Also, over the range of the 2.2, the time averaged value of the mean drag coefficient increases with the increasing value of the power-law index (n), i.e., as the fluids behavior changes from shear-thinning (n<1) to Newtonian (n=1) and then to shear-thickening (n>1).The trend of variation of the mean drag coefficient is reversed for 2.2. Due to this change, it is obvious that there should be some points in which the mean drag coefficient is independent of the power-law index (i.e. in a same rotational speed, the mean drag coefficient does not change with n). Variation of the mean lift coefficient against dimensionless rotational speed () at different power-law index (n) is illustrated in Figure 14. For all the cases considered power-law index (n), mean lift coefficient augments with increasing dimensionless rotational speed (). Because of the rotation of counter-clockwise, greater momentum acts on the upper part of the cylinder and causes the lift force orients toward the negative direction. Increasing the rotation, makes the difference between the momentum acts on the upper and bottom part of the cylinder greater and so the absolute value of the lift coefficient gets bigger. At the same dimensionless rotational speed (), the diagram shows that the absolute mean lift coefficient augments with increasing power-law index (n). Also, It can be seen the increment of the mean lift coefficient is governed by the linear function of CL = m., where m is the slope. The values of slopes for different power-law index (n) are presented in Table 3. Figure 15 shows the variation of Strouhal number with dimensionless rotational speed ratio at different power-law index (n). It can be seen that the variation of Strouhal number can be divided into two major parts. For n<1, the value of Strouhal number increases slightly at low dimensionless rotational speed ratios (<1), and then decreases with increasing . While for
18

n1 the diagram shows a decreasing trend with increasing . For a fixed value of dimensionless rotational speed ratio, the value of Strouhal number decreases with an increasing value of power-law index (n). Also, according to figure 15, the effect of powerlaw index (n) on Strouhal number is overall very strong compared with rotational speed.

4.4. Pressure coefficient ( CP )


It is generally believed that, in flow over a bluff body such as a circular cylinder and a sphere, the effect of pressure on drag and lift forces is more significant than that of friction. Therefore, pressure distribution around the cylinder surface is needed to further understand the effect of rotation and power-law index (n) on the lift force. Figure 16 depicts mean pressure coefficient according to the power-law index (n) for stationary cylinder. It can be seen that for no rotational speed, pressure distribution for different power-law index (n) is almost symmetric (=180) leading to a zero mean lift. CP is maximum in the forward stagnation point (=180), then it decreases along the surface toward the rear of the cylinder (=360). The minimum value of the pressure coefficient occurs at the point of separation (=S 270)). With increasing , CP rises again at the back of the cylinder. This is obviously because of the recirculation in the rear of the cylinder. Figure 17 shows the time-averaged pressure coefficient CP as a function of circumferential direction for different rotational speeds at three different power-law index (n). As increases the diagram becomes asymmetric and the location of minimum pressure coefficient moves to lower side for a fixed value of the flow behavior index (n). This aspect can explain augmentation of the absolute lift coefficient. Compared to the power-law index (n), the results exhibit that rotational speed has more significant influence on the pressure coefficient

19

at the cylinders surface. Also, with increasing n, the minimum value of the pressure coefficient increases in its magnitude.

5. Conclusions
In this study, non-Newtonian flow passing a rotating circular cylinder in the 2D laminar flow regime has been investigated numerically usingthe Lattice Boltzmann Method with multirelaxation-time (MRT) collision model. The study was carried out at a various range of power-law index (0.4n1.8) and dimensionless rotational ratio (02.5) at Reynolds number of 100. The numerical simulations were validated against the experimental and numerical results. Conclusion of the present study can be summarized as follows: (i) It was found that LBM is a suitable approach for the simulation of the non-Newtonian flows in the geometries which include curved boundaries. In addition, the viscous stresses can be calculated locally that is the most preference of this method. (ii) In the case of no rotation, = 0, it is seen that as n increases the length and width of the wake increases. (iii) For a fixed value of the power-law index (n), with increasing dimensionless rotational speed (), all the vortices become narrower compared to the case of no rotation, leading to decrease in the lateral width of wake. Furthermore, vortex shedding pattern disappears beyond a certain threshold value and the flow is stabilized. (iv) The mean drag coefficient decreases with the increasing dimensionless rotational speed at the same power-law index (n). The maximum decrease in the drag values is seen for shear-thickening (n>1), especially for n=1.8. Also, over the range of the 2.2, the time averaged value of the mean drag coefficient increases with the increasing value of the power-law index (n), i.e., as the fluid behavior changes from shear-thinning (n<1) fluids
20

to Newtonian (n=1) and then to shear-thickening (n>1).The trend of variation of the mean drag coefficient is reversed for 2.2. Due to this change, it is obvious that there should be some points in which the mean drag coefficient is independent of the power-law index (i.e. in a same rotational speed, the mean drag coefficient does not change with n). (v) For all the considered power-law indices (n), mean lift coefficient augments with increasing dimensionless rotational speed (). Also, at the same dimensionless rotational speed (), the diagram shows that the absolute mean lift coefficient augments with increasing power-law index (n). (vi) Variation of Strouhal number with dimensionless rotational speed ratio at different power-law index (n) can be divided into two major parts. For n<1, the value of Strouhal number increases slightly at low dimensionless rotational speed ratios (<1), and then decreases with increasing . While for n1, the value of Strouhal number decreases with increasing . Generally, the effect of power-law index (n) on Strouhal number is overall very strong in comparison with rotational speed.

References:
[1] C. H. K. Williamson, Oblique and parallel modes of vortex shedding in the wake of a circular cylinder at low Reynolds number, J. Fluid Mech. 206 (1989) 579627. [2] M. Zdravkovich, Flow around circular cylinders, vol. 1, Fundamentals, Oxford University Press, New York, 1997.

21

[3] M. Coutanceau, C. Menard, Influence of rotation on the near-wake development behind an impulsively started circular cylinder, J. Fluid Mech. 158 (1985) 399466. [4] Y.Y. Yan, Y.Q. Zu, Numerical simulation of heat transfer and fluid flow past a rotating isothermal cylinder A LBM approach, Int. J. Heat Mass Transfer. 51 (2008) 25192536. [5] S. Mittal, B. Kumar, Flow past a Rotating Cylinder, J. Fluid Mech. 476 (2003) 303-334. [6] S. Kang, H. Choi, S. Lee, Laminar flow past a rotating circular cylinder, Physics of Fluids 11 (1999) 3312-3321. [7] P.M.Coelho, F.T. Pinho, Vortex shedding in cylinder flow of shear-thinning fluids. I. Identification and demarcation of flow regimes, J. Non-Newtonian Fluid Mech.110 (2003) 143176. [8] P.M.Coelho, F.T. Pinho, Vortex shedding in cylinder flow of shear-thinning fluids, II. Flow characteristics, J. Non-Newtonian Fluid Mech. 110 (2003) 177193. [9] P.M. Coelho, F.T. Pinho, Vortex shedding in cylinder flow of shear-thinning fluids. III. Pressure measurements, J. Non-Newtonian Fluid Mech. 121 (2004) 5568. [10] P. Sivakumar, R. P. Bharti, R.P. Chhabra, Effect of power-law index on critical parameters for power-law flow across an unconfined circular cylinder, Chem. Eng. Sci. 61 (2006) 6035 6046. [11] R. P. Bharti, R. P. Chhabra, V. Eswaran, Steady flow of power-law fluids across a circular cylinder. Can J Chem Eng 84 (2006) 406421. [12] R. C. Patil, R. P. Bharti, R. P. Chhabra, Steady Flow of Power Law Fluids over a Pair of Cylinders in Tandem Arrangement, Ind. Eng. Chem. Res., 47 (2008) 1660 1683. [13] P. Sivakumar, R.P. Bharti, R.P. Chhabra, Steady flow of power-law fluids across an unconfined elliptical cylinder, Chem. Eng. Sci. 62 (2007) 16821702. [14] V.K. Patnana, R.P. Bharti, R.P. Chhabra, Two-dimensional unsteady flow of power-law fluids over a cylinder, Chem. Eng. Sci. 64 (2009) 29782999.
22

[15] Saroj K. Panda, R.P. Chhabra, Laminar flow of power-law fluids past a rotating cylinder, J. Non-Newtonian Fluid Mech. 165 (2010) 14421461. [16] T. Cochrane, K. Walters, M.F. Webster, On Newtonian and non-Newtonian flow in complex geometries, Phil. Trans. Soc. Lond. A., 301 (1981) 163-181. [17] S. Chen, H. Chen, D.O. Martinez, W.H. Matthaeus, Lattice Boltzmann model for simulation of magnetohydrodynamics, Phys. Rev. Lett. 67(27) (1991) 3776-3779. [18] Y.H. Qian, D. d'Humires, P. Lallemand, Lattice BGK model for Navier_Stokes equation, Europhys. Lett. 17 (1992) 479-485. [19] S. Chen, G.D. Doolen, Lattice Boltzmann method for fluid flow, Annu. Rev. Fluid Mech. 30 (1998) 329-364. [20] E. Fattahi, M. Farhadi, K. Sedighi, Lattice Boltzmann simulation of natural convection heat transfer in eccentric annulus, Int. J. Therm. Sci. 49(12) (2010) 2353-2362. [21] GH.R. Kefayati, S.F. Hosseinizadeh, M. Gorji, H. Sajjadi, Lattice Boltzmann simulation of natural convection in tall enclosures using water/SiO2 nanofluid, Int. Comun. Heat. Mass. 38(6) (2011) 798-805. [22] H., Nemati, M., Farhadi, K., Sedighi, E., Fattahi, A.A.R, Darzi, Lattice Boltzmann simulation of nanofluid in lid-driven cavity, Int. Comun. Heat. Mass. 37(10) (2010) 15281534. [23] D. Gao, Z. Chen, Lattice Boltzmann simulation of natural convection dominated melting in a rectangular cavity filled with porous media, Int. J. Thermal Sci. 50(4) (2011) 493-501. [24] E. S. Boek, J. Chin, P. V. Coveney, Lattice Boltzmann simulation of the flow of nonNewtonian fluids in porous media, Int. J. Mod Phys. C. 17 (2003) 99-102. [25] C. Pan, L.S. Luo, C.T. Miller, An evaluation of lattice Boltzmann schemes for porous medium flow simulation, Comput. Fluids 35 (2006) 898-909.

23

[26] E.S. Boek, M. Venturoli, Lattice-Boltzmann studies of fluid flow in porous media with realistic rock geometries, Comp. Math. Appl., 59(2010) 2305-2314. [27] A. Kuzmin, A. A, Mohamad, S. Succi, Multi-Relaxtion Time Lattice Boltzmann model for multi phase flows, Int. J. Mod Phys. C. 9(6) (2008) 875-902. [28] H. Nemati, M. Farhadi, K. Sedighi, M.M. Pirouz, E. Fattahi, Numerical simulation of fluid flow around two rotating side by side circular cylinders by Lattice Boltzmann method, Int. J. Comput. Fluid Dyn. 24(3) (2010) 83-94. [29] K. Fallah, A. Fardad, E. Fattahi, N. Sedaghati zadeh, A. Ghaderi, Numerical simulation of planar shear flow passing a rotating cylinder at low Reynolds numbers, Acta Mech 223 (2012) 221236. [30] K. Fallah, A. A. Fardad, E. Fattahi, N. Sedaghati Zadeh, Simulation of Planar Shear Flow Passing Two Equal Sized Circular Cylinders in Tandem Arrangement, AIP Conf. Proc. 1400 (2011) 449-452. [31] S. Gabbanelli, G. Drazer, J. Koplik, Lattice Boltzmann method for non-Newtonian (Power-law) fluids, Phys. Rev. E. 72 (2005) 046312. [32] J. Boyd, J. Buick, S. Green, A second-order accurate lattice Boltzmann non- Newtonian flow model, J. Phys. A: Math. Gen. 39 (2006) 1424114247. [33] O. Malaspinas, G. Courbebaisse, M. Deville, Simulation of generalized Newtonian fluids with the lattice Boltzmann method, Int. J. Mod. Phys. C 18 (2007)19391949. [34] S.P. Sullivan, L.F. Gladden, M.L. Johns, Simulation of power-law fluid flow through porous media using lattice Boltzmann techniques, J. Non-Newton. Fluid Mech. 133 (2006) 9198. [35] M. Yoshino, Y. Hotta, T. Hirozane, M. Endo, A numerical method for incompressible non-Newtonian fluid flows based on the lattice Boltzmann method, J. Non-Newton. Fluid Mech. 147 (2007) 6978.
24

[36] C.H. Wang, J.R. Ho, Lattice Boltzmann modeling of Bingham plastics, Physics A. 387 (2008) 47404748. [37] A. Vikhansky, Lattice-Boltzmann method for yield-stress liquids, J. Non-Newton. Fluid Mech. 155 (2008) 95100. [38] G.H. Tang, X.F. Li, Y.L. He, W.Q. Tao, Electroosmotic flow of non-Newtonian fluid in microchannels, J. Non-Newton. Fluid Mech. 157 (2009) 133137. [39] M. Ashrafizaadeh, H. Bakhshaei, A comparison of non-Newtonian models for lattice Boltzmann blood flow simulations, Comput. Math. Appl. 58 (2009) 10451054. [40] M.M. Hedayat, M.H. Borghei, A. Fakhari, K. Sadeghy, On the Use of LatticeBoltzmann Model for Simulating Lid-Driven Cavity Flows of Strain-hardening Fluids, Nihon Reoroji Gakkaishi (J. Soc. Rheol. Jpn.), 38 (2010) 201-207. [41] S. Ubertini, S. Succi, Recent advances of Lattice Boltzmann techniques on unstructured grids, Prog. Comput. Fluid Dyn. 5(2005) 8596. [42] P.L. Bhatnagar, E.P. Gross, M. Krook, A model for collision processes in gases. I. Small amplitude processes in charged and neutral one-component systems, Phys. Rev. 94 (1954) 511525. [43] P. Lallemand, L. S. Luo, Theory of the lattice Boltzmann method: dispersion, dissipation, isotropy, Galilean invariance, and stability, Phys. Rev. E. 61 (2000) 65466562. [44] I. Ginzburg, D. dHumieres, Multireflection boundary conditions for lattice Boltzmann models, Phys. Rev. E. 68 (2003) 066614. [45] D. dHumieres, Generalized lattice-Boltzmann equations, in Rarefied Gas Dynamics: Theory and Simulations, Prog. Aeronaut. Astronaut. 159 (1992) 450458. [46] Z. Chaia,, B. Shia, Z. Guo, , F. Rong, Multiple-relaxation-time lattice Boltzmann model for generalized Newtonian fluid flows, J. Non-Newton. Fluid Mech. 166 (2011), 332-342.

25

[47] O. Filippova, D. Hnel, Grid refinement for lattice-BGK models, J. Comput. Phys. 147 (1998) 219-228. [48] R. Mei, L-S. Lou, W. Shyy, An accurate curved boundary treatment in the lattice Boltzmann method, J. Comput. Phys. 155 (1999) 307-330. [49] P. Lallemand, L.S. Luo, Lattice Boltzmann method for moving boundaries, J. Comput. Phys. 184 (2003) 406 421. [50] Z.L. Guo, Ch. Zheng, B.C. Shi, An extrapolation method for boundary conditions in lattice Boltzmann method, Phys. Fluids. 14 (2002) 20072010. [51] Q. Zou, X. He, On pressure and velocity flow boundary conditions for the lattice Boltzmann BGK model, Phys. Fluids. 9 (1997) 15911598. [52] D. Yu, R. Mei, W. Shyy, Improved treatment of the open boundary in the method of lattice Boltzmann equation, Prog comput fluid dy. 5 (2005) 111. [53] R. Mei, L-S. Lou, W. Shyy, Force evaluation in the lattice Boltzmann method involving curved geometry, Phys. Rev. E. 155 (2002) 307-330.

26

Figures
Figure1.DiscretevelocityvectorsfortheD2Q9model Figure2.Figureoflatticesincurvedwallboundary. Figure3.Bouncebackfordifferent . Figure4.Flowgeometry,coordinatesystemandcomputationaldomain. Figure5.ComparisonbetweentheevolutionofthevelocityfieldforRe=200,=0.5.(Left:obtainedbypresent computation;middle:obtainedbynumericstudy(YanandZu[4]);right:obtainedbyexperiment(Coutanceau andMenard[3])). Figure 6.Patterns of vorticity contour for different power law index at Re=100 and =0. Figure 7.Patterns of vorticity contour for different speed ratios (a) n=0.4, (b) n=1.0, and (c) n=1.8. Figure 8.Patterns of vorticity contour for bigger (a) n=1.0, (b) n=1.4. Figure 9.Time evolutions of the drag and lift coefficient for different power-law index at Re = 100 and =0. Figure 10.Time evolutions of the drag and lift coefficient for different dimensionless rotational speed () for n=0.4 . Figure 11.Time evolutions of the drag and lift coefficient for different dimensionless rotational speed () for n=1. Figure 12.Time evolutions of the drag and lift coefficient for different dimensionless rotational speed () for n=1.8. Figure 13.Mean drag coefficients versus dimensionless rotational speed () for different power-law index (n). Figure 14.Mean lift coefficients versus dimensionless rotational speed () for different power-law index (n). Figure 15.Strouhal number (St) versus dimensionless rotational speed () for different power-law index (n). Figure16.Meanpressurecoefficientsaroundthecylinderatdifferentpowerlawindex(n) Figure17.Meanpressurecoefficientsaroundthecylinderplottedaccordingtoatdifferentpowerlawindex (n)(a:n=0.4;b:n=1.0;c:n=1.8).

27

28

Table 1 Comparison between time-averaged drag coefficient (C D ) , time-averaged lift coefficients (C L ) and Strouhal number (St) with Yan and Zu [4] for different dimensionless rotational ratio (), at Re = 200.

Mesh

CD
1.587 1.553 1.553 1.505 1.478 1.432 1.394 1.372 1.349

CL
0.2698 0.2669 1.384 1.331 2.809 2.771 2.744 2.728 2.699

St

PresentStudy 0.1 YanandZu[4] PresentStudy 0.5 YanandZu[4]

1200400 1200400 600200 900300

0.1084 0.1089 0.1099 0.1096 0.1231 0.1164 0.1110 0.1111 0.1100

PresentStudy 1.0 1200400 1500500 YanandZu[4]

29

Table2ComparisonofthesteadyandunsteadynonNewtonianflowcharacteristicsofastationarycircular cylinder.

Re

CD
1.2621 1.2321 1.2459 1.3775 1.3368 1.3576 1.551 1.5350

St

PresentStudy 0.4 Sivakumaretal.[13] Patnanaetal.[14] PresentStudy 0.6 Sivakumaretal.[13] Patnanaetal.[14] Steady 40 PresentStudy 1.0 Sivakumaretal.[13] Patnanaetal.[14]

PresentStudy 1.4 Sivakumaretal.[13] Patnanaetal.[14]

1.678 1.6061 1.6541

Unsteady

140 0.6

PresentStudy Patnanaetal.[14] PresentStudy 1.0 Patnanaetal.[14]

1.2385 1.2236 1.3410 1.3281

0.2141 0.2066 0.1857 0.1800

30

PresentStudy 1.8 Patnanaetal.[14]

1.6101 1.5868

0.1571 0.1511

31

Table3Relationshipbetweenpowerlawindex(n)andslope(m). n m 0.4 2.2297 0.6 2.4772 0.8 2.7453 1 2.9842 1.2 3.1963 1.4 3.4003 1.6 3.6072 1.8 3.8141

32


Figure1

33

Figure2

34

(a)

(b)

(c)

Figure3

35

Figure4

36

(a) t*=3

(b) t*=7

(c) t*=9

(d) t*=11

Figure5

37

(a)n=0.4 (b)n=0.6

(c)n=0.8 (d)n=1.0

(e)n=1.2 (f)n=1.4

(g)n=1.6 Figure 6 (h)n=1.8

38

= 0.5

= 1.0

= 1.5

= 2.0 = 2.5 (a) n=0.4 (b) n=1.0 Figure 7 (c) n=1.8

39

= 3

= 4

= 5

(a) n=1.0 Figure 8 (b) n=1.4

40

Figure 9

41

Figure 10

42

Figure 11

43

Figure 12

44

Figure 13

45

Figure 14

46

Figure 15

47


Figure16

48

Figure17

49

The Power-law fluid flow around a rotating cylinder is studied numerically. Lattice Boltzmann Method with multi-relaxation-time collision model is used. Increasing the rotational speed will decrease drag coefficient. Increasing the power-law index will increase length and width of the wake. Vortex shedding pattern disappears beyond a threshold rotational speed.

50

You might also like