Acta Mecanica manuscript No. (will be inserted by the editor)

Ozan Tu˘gluk · Hakan I. Tarman

Direct Numerical Simulation of Pipe Flow Using a Solenoidal Spectral Method

the date of receipt and acceptance should be inserted later

Abstract In this study, a numerical method based on solenoidal basis functions, for the simulation of incompressible flow through a circular-cylindrical pipe is presented. The solenoidal bases utilized in the study are formulated using the Legendre polynomials. Legendre polynomials are favorable, both for form of the basis functions, and for the inner product integrals arising from the Galerkin type projection used. The projection is performed onto the dual solenoidal bases, eliminating the pressure variable, simplifying the numerical approach to the problem. The success of the scheme in calculating turbulence statistics, and its energy conserving properties are investigated. The generated numerical method is also tested by simulating the effect of drag reduction due to spanwise wall oscillations.

1 Introduction Turbulent flow in a pipe has been an intriguing topic for research for well over a century, ever since the famous experiments by Reynolds [23]. In those experiments, Reynolds observed how direct flow in a pipe became sinuous. This phenomenon, which is now called transition to turbulence, has been subject to diverse numerical, theoretical and experimental studies. It is now widely accepted that, the sole parameter of this phenomenon is the Reynolds number. Even though the pipe flow is geometrically very simple, and experiments are relatively easy to conduct in this setting, the theoretical studies were limited, and the numerical approach was more widely used. One of the main peculiarities of pipe flow is that the flow is linearly stable at even very high Reynolds numbers [15], causing a decay in small perturbations. This means, for transition in pipe flow to occur, both the flow speed and the perturbation amplitude must be sufficiently high [5]. Ozan Tu˘gluk · Hakan I. Tarman Department of Engineering Sciences, METU, [email protected], [email protected]

06531

Ankara/TURKEY

E-mail:

2

As mentioned earlier, investigations of pipe flow generally utilize direct numerical simulations (DNS). These DNS are usually implemented via spectral (pseudospectral [24], spectral element [4], etc.) or finite difference/finite volume methods [6,8], or a mixture of these method families. Within the framework of spectral methods, it is possible to employ solenoidal basis functions in the expansion of velocity. The use of solenoidal functions in fluid mechanics problems appeared earlier in various studies such as [17] in the study of Taylor-Couette flow, and in the study of pipe flow [12,15]. The convergence properties of expansions in divergence-free functions in the case of Stokes equations along with pressure reconstruction algorithm is studied in [19], and a domain decomposition approach can be found in [7]. Another early formulation of the incompressible fluid flow problems in terms of divergence-free functions can also be found in [16]. Solenoidal basis functions are vector functions with zero divergence, and they span a divergence-free space. When one uses these basis functions to expand the velocity field in the incompressible flow problems, the continuity equation is automatically satisfied. In addition, Hermitian projection over a solenoidal field cancels out any gradient term and therefore no pressure solvers are required, resulting in simpler computer code. Furthermore the solution is guaranteed to be divergence free, even for low resolutions. Overall, the solenoidal basis functions provide a simple and flexible framework for the numerical analysis of structures in the transition regime which is beneficial for the understanding of the underlying physics of transition and for the attempts at its control in incompressible flow. The rest of this work is organized as follows, the governing equations and the scalings used are given in Section 2, details about our numerical method is presented in Section 3, turbulence statistics at a reference Reynolds number are presented in Section 4, along with the simulation of drag reduction via spanwise wall oscillations. Finally Section 6 summarizes our findings and our comments on the strengths and weaknesses of the numerical method.

2 Governing Equations The governing system of unsteady incompressible Navier-Stokes (N-S) equations in polar coordinates, when non-dimensionalized with a reference velocity ure f and reference length lre f , can be written as,

∂t u + (u · ∇)u = −∇p + ∇ · u = 0,

1 ∆u Reref

(1)

where, u = (u, v, w), denotes the velocity field with radial, azimuthal and axial components, respectively and p = p′ + Gz the pressure. In the given equations Rere f = ure f lre f /ν represents the Reynolds number, with kinematic viscosity ν , and G is the non-dimensional magnitude of the mean pressure gradient along the pipe axis, ez . They are stated below for various scalings employed in literature. When the velocity scale is chosen as the centerline velocity, Ucl , of laminar flow, and the length scale is the pipe radius R (also called as macro scaling, and

3

the resulting units are called macro units), Recl =

Π R3 ρ 4µ 2

and

G=

4 , Recl

(2)

where Π is the dimensional driving pressure gradient, ρ the density and µ is the dynamic viscosity. On the other hand, when the velocity scale is the friction velocity and the length scale is again the pipe radius (also called friction or inner scaling and the resulting units are called friction units or inner units), Reτ =

uτ R ν

and G = 2,

(3)

p where uτ = τw /ρ is the wall friction velocity and τw the wall shear stress. In these scalings, the mean (driving) pressure gradient is fixed. Another scaling arises from fixing the bulk mean velocity, UB =

1 π

Z 2π Z 1 o

0

wr drd θ ,

and letting the pressure gradient vary in time. In this case velocity is scaled by 2 UB , which is the centerline velocity of the corresponding laminar flow. In this scaling the Reynolds number is defined as , Re =

2UB R ν

and G = G(t),

(4)

where G(t) is the time varying pressure gradient required to keep the bulk velocity constant. In this study the constant pressure gradient and constant mass flux (constant bulk mean velocity) cases will be investigated. The solution domain is r ∈ (0, 1],

θ ∈ [0, 2π ),

z ∈ [0, L).

In this domain, the flow is naturally periodic along azimuthal (θ ) direction and taken to be periodic along the axial (z) direction. The no-slip condition u(1, θ , z,t) = 0

(5)

is imposed at the pipe wall. Under the scaling in (2), laminar base flow becomes uB = (0, 0, 1 − r2 ). The present numerical approach for solving (1) is essentially a Galerkin type scheme involving the solenoidal basis functions satisfying the boundary conditions for the expansion of the velocity field and the solenoidal dual basis functions for the construction of the projection space. In the process, the pressure variable in (1) is eliminated. The next section details this approach.

4

(a) Clustered grid

(b) Non-clustered grid

Fig. 1 Comparison of clustered and non-clustered grids for 8 radial grid points

3 Numerical Method A similar approach utilizing the solenoidal basis in the numerical study of the pipe flow are detailed in [14,15], where Chebyshev polynomials are used for the construction of the bases. In this work, however, Legendre polynomials are used in the formulation of the solenoidal basis functions and their duals. This is preferred for their performance in the numerical evaluation of the inner product integrals and for the resulting form of the solenoidal basis functions. We employ two families of basis functions. The physical (trial) basis functions are used to expand the velocity field, whereas the dual (test) basis functions are used for the projection purposes. In summary, the numerical method consists of expanding the velocity field in terms of physical basis functions, substituting into the governing equations and projecting the resulting residual equations onto the dual bases for the elimination of the residual error. After the projection, space dependence in the problem and the pressure term are eliminated, yielding a dynamical system for the time dependent expansion coefficients. For spatial discretization, equidistant grids are used along the periodic directions, Gauss-Legendre grid is used along the radial direction. Special attention is paid to the grid distribution in the vicinity of the pipe center r = 0, in order to avoid the singularity and to prevent grid clustering [25]. The grid clustering phenomena in polar geometries is illustrated in Fig. 1. 3.1 Basis Functions There are two sets of conditions to be satisfied by the physical Ψ and the dual Φ basis function families. The first one is the solenoidal condition, i.e. the basis functions should be divergence free, ∇ ·Ψ = 0 ∇ · Φ = 0.

(6)

5

It is worthwhile to note that the above formulation for the dual bases is only valid, if the dual bases are constructed using Legendre polynomials, which are orthogonal with respect to weight unity. If for example Chebyshev polynomials are used, √ as in [15], the condition becomes ∇ · (ω Φ ) = 0, where ω = 1/( 1 − r2 ), which further complicates the construction of the solenoidal bases. In addition to the above condition, the physical basis functions are required to satisfy the no-slip boundary conditions prescribed in (5). On the other hand, the dual bases are only required the satisfy the condition of the vanishing flux through the walls in order to enable the elimination of the pressure term ∇p′ from the equations [17]. These additional conditions are

Ψ (1, θ , z) = 0 Φ (1, θ , z) · er = 0.

(7)

In the pipe geometry, while the flow in the azimuthal direction is naturally periodic, the flow in the axial direction is taken to be periodic as it is a common practice in literature. This leads to the use of Fourier representation along the corresponding θ , z coordinates and the basis functions take the following form,

Ψlnm (r, θ , z) = ei(nθ +2π lz/Q) Vlnm (r) (1,2)

(1,2)

¯ Φlnm (r, θ , z) = ei(nθ +2π lz/Q) V lnm (r). (1,2)

(1,2)

The superscripts on the basis functions signify the sufficiency of the two degrees of freedom in representing the three components of a solenoidal velocity field as the continuity equation provides the connection between the components. In the ¯ are required Fourier representation, the three components (Vr ,Vθ ,Vz ) of V and V to satisfy the reduced form of the continuity equation, D+Vr +

in Vθ + ilVz = 0, r

(8)

where D+ = D + 1r and D = d/dr. The basis functions are then constructed to satisfy (7) and (8). The regularity requirement in the vicinity of the pipe center [20, 13] and the use of Gauss-Legendre quadrature (ωk , rk ) in the numerical evaluation of the inner product integrals,

¯ V) = (V,

Z1 0

¯ ∗ · V(r) rdr = 1 V(r) 2

Z1

−1

¯ ∗ · V(r) rdr V(r)

1 K ¯ ∗ = ∑ V(r k ) · V(rk ) rk ωk , 2 k=0

(9)

impose additional requirements on the basis functions such as the evenness of the integrands. The details are given in Appendix A.

6

3.2 Projection Having constructed the basis functions, the projection procedure is now employed to reduce (1) to a dynamical system via the inner product (Ψ , Φ ) =

Z1

rdr

0

Z2π



ZL 0

0

dz(Ψ ∗ · Φ )

(10)

which reduces to (9) due to the orthogonality of Fourier bases. The coefficients a(1) , a(2) in the expansion of the velocity field u(r, θ , z,t) =

Q

N

M

∑ ∑ ∑

(1)

(2)

(1)

(2)

alnm (t) Φlnm (r, θ , z) + alnm (t) Φlnm (r, θ , z)

(11)

l=−Q n=−N m=0

can then be obtained by solving " (1) # " (1) # " (1) # ¯ ′ , uˆ ln ) ¯ (1) ′ , V(2) ) ¯ ′ , V(1) ) (V (V a˙lnm ∂ (V lnm lnm lnm lnm lnm Alnm′ m a˙ lnm ≡ = (2) ∂ t (V ¯ (2) ′ , uˆ ln ) ¯ (2) ′ , V(1) ) (V ¯ (2) ′ , V(2) ) (V a ˙ lnm lnm lnm lnm lnm lnm for each wave number pair (l, n) and for K = 3M in the least to render the numerical quadrature exact in the nonlinear case, where u(r, θ , z,t) =

Q

N

∑ ∑

ei(nθ +2π lz/L) uˆ ln (r,t).

l=−Q n=−N

The substitution of (11) into (1), and the projection of the resulting residual onto the dual bases via (10) result in the dynamical system for the expansion coefficients alnm in the form Alnm′ m a˙ lnm = Blnm′ m alnm − blnm′

(12)

where, 1 ∆ Φlnm − ∇p) Re = (Ψlnm′ , (u · ∇)u).

Blnm′ m = (Ψlnm′ , blnm′

The pressure term ∇p appears in the expression, but the term ∇p′ cancels under projection. The system (12) is numerically integrated in time using the 3rd order semi-implicit time-solver, based on backward difference and Adams-Bashford methods, (11A − 6∆ t B) a(k+1) = A(18a(k) − 9a(k−1) + 2a(k−2) ) − ∆ t(18b

(k)

− 18b

(k−1)

+ 6b

(k−2)

(13) ),

for the expansion coefficients a, and they are subsequently used to construct the velocity field (11) and the projection b for the computation of the next step. This scheme (13) is not self starting, so for the initial steps we use a Runge-Kutta time integrator of corresponding order.

7

M 20 30 40 50 50 [13]

λ1 -0.022998556376 + 0.950174903582i -0.023170716674 + 0.950481315400i -0.023170795763 + 0.950481396669i -0.023170795764 + 0.950481396669i -0.023170795764 + 0.950481396670i

Table 1 Leading eigenvalue of (14) corresponding to Recl = 9600, and (l=1, n=1)

4 Benchmark case for turbulent pipe flow at low Reynolds number As an initial means of testing our method, we consider the linearization of (1) about the laminar base flow, uB , and assume an exponential time evolution, eλ t , of the velocity. These operations and assumptions result in an independent generalized eigenvalue problem, for each wave-number pair (l, n). (λ Alnm′ m − Blnm′ m (Re))alnm = 0.

(14)

We compute the eigenvalues, λ , of (14), at the reference Reynolds number Recl = 9600. The results are given in Table 1. The agreement to 10 digits is obtained with those computed and cited in [15]. It is also possible to observe the convergence behaviour from Table 1, as the resolution is increased. After completing the validation for the linear case, we compute the turbulence statistics at Reτ = 180. The computations were performed for a pipe with a length of 10 radii (i.e L = 10). The results are presented in Table 2 and in Figures 2-5, in comparison with the DNS and experimental results presented in [6], and the numerical results from [8]. Fukagata et al. [8] use a highly energy conserving scheme based on a finite difference method, with a resolution (radial × azimuthal × axial) of 96 × 128 × 256, whereas Eggels et al. [6] use a finite volume discretization with the same resolution. In our method we do not impose any explicit energy conserving constraint on the system and use an effective resolution of 49 × 81 × 151. Our computations are fully de-aliased in the periodic directions using the 3/2 rule and, a time step of 2.5 × 10−4 non-dimensional time units. The statistics are computed using 80 flowfields equispaced in time with a spacing of 0.25 non-dimensional time units after discarding the initial transients. It is worthwhile to note that even though we use a coarser grid, we have seven grid points within the viscous sublayer whereas [6] have only three. Thus, our grid is denser near the walls in comparison. It is widely known that spectral methods have superior energy conservation properties [1,9,25,11] when compared to finite difference methods. In addition Galerkin type projections are known to be stable in the energy sense [11]. In regards to turbulence statistics, in Figures (2-5); it can observed that, our method can predict the turbulence statistics with a relative error in the order of 3%. The main difference with the reference calculations is that, our code predicts rms velocity fluctuations closer to the experimental results [6] in the near wall region (Figure 4) and seems to match the peak experimental Reynolds stress magnitude closely (Figure 5). We believe that this increased predictive power at a lower resolution comes from the combination of accuracy of spectral methods, inherent satisfaction of the continuity equation by solenoidal bases expansion and the fact that we have a denser grid composed of Gauss-Legendre points in the near

8

Ucl /UB UB /uτ Ucl /uτ cf Recl Re Reτ

Present (1) 1.32 14.58 19.27 0.00941 6864 5200 180

Present (2) 1.31 14.08 18.45 0.01008 6419 4900 174

Quadrio et al. [22] (DNS) (2) 1.31 14.24 18.63 0.00963 6400 4900 172

Eggels et al.[6] (DNS) (1) 1.31 14.73 19.31 0.00922 6950 5300 180

Eggels et al. [6] (PIV) (1) 1.30 14.88 19.38 0.00903 7100 5450 183

Table 2 Turbulence statistics for pipe flow in the cases of (1) constant pressure gradient, (2) constant mass flux. N 25 50 100

∆ t, clustered grid 0.00282 0.000775 0.0000477

∆ t, non-clustered grid 0.0156 0.00625 0.00196

Table 3 Maximum allowable time step, in units of 2t UB /R

wall region. However, when comparing the results for the core region (near the pipe centerline), our method under predicts the rms fluctuations of the azimuthal velocity by a fair margin. In addition, the computed mean axial velocity is lower than suggested by experimental results. This however, is expected as our grid is relatively coarse in the core region. In the light of these observations it is safe to say that, at the resolutions employed, our method provides increased accuracy in the wall region at the expense of predictive power in the center region. In addition to friction scaling, a simulation keeping the bulk velocity constant is also performed, at Re = 4900. This Reynolds number is chosen to facilitate comparison with [22], and is about 6% lower than the bulk Re calculated in our earlier simulation at Reτ = 180. In keeping the bulk velocity constant, at a value of 0.5 after non-dimensionalization, an adjustable pressure gradient is used in contrast to a constant pressure gradient used in the friction scaling case. For this case the simulation has been carried out for 600 non-dimensional time units after discarding the initial transients as in [22], a total of 150 flowfields equidistant in time were employed in calculation of the statistics. The findings are summarized in Table 2 along with the earlier studies ([6],[22]). During our computations, the average CFL number computed to be equal to 0.4, as defined by, CFL = max



urms ∆ t ∆r

 ∆t wrms ∆ t , + vrms + r∆ θ ∆z

over the grid, where the subscript rms denotes root mean square value of velocity component. As mentioned before, our code uses a non-clustered grid configuration in the radial direction (see Fig.1). Table 3 gives the maximum allowable time step (CFL < 1), as a function of number of radial grid points for both clustered and non clustered cases for the reference velocities computed above, it is evident that the time step restriction is quite severe for a clustered grid configuration. The code, with the resolution stated above, uses approximately 4 Gb of memory and serial calculation for a single time step takes 2.1 seconds on a 2.8 GHz i7 CPU.

9

1

20 18 16 14 12

wm 10 8 6 4 2 0

1

10

100

r+ Fig. 2 Semi-logarithmic plot of mean axial velocity,’◦’ PIV Eggels et al.[6], ’’ Fukagata et al. [8], ’♦’ present study where r+ = (1 − r)Reτ

1

0.8

wm Ucl

0.6

0.4

0.2

0

0

20

40

60

80

100

120

140

160

180

r+ Fig. 3 Mean axial velocity scaled by mean centerline velocity, ’◦’ PIV Eggels et al.[6], ’’ Fukagata et al. [8], ’♦’ present study.

10

3

2.5

w′

2

u′ , v′ , w′ v′

1

u′ 0.2

0

0

20

40

60

80

100

120

140

160

180

r+ Fig. 4 Rms velocity fluctuations,’◦’ PIV Eggels et al.[6], ’’ Fukagata et al. [8], ’♦’ present study. 0.8

0.7

0.6

0.2

u′ w′

0.4

0.3

0.2

0.1

0

0

20

40

60

80

100

120

140

160

180

r+ Fig. 5 Mean Reynolds stress, u′ w′ , ’◦’ PIV Eggels et al.[6], ’’ Fukagata et al. [8], ’♦’ present study.

11

Resolution, time step Current, 13 × 16 × 32 ∆ t = 0.01 Current, 13 × 16 × 32 ∆ t = 0.001 Fukagata et. al. [8], 12 × 16 × 32 ∆ t = 0.0001

t1 470.8 471.1 2.71

t50 799.9 800.1 34.16

tdiv 856.1 856.4 167.13

Table 4 Growth of kinetic energy as Reτ → ∞

The energy conserving property of our method as Reτ → ∞ is also investigated. For this investigation we take an instantaneous velocity field from our DNS calculations and interpolate it on a much coarser grid, ensuring that the mean velocity is equal to zero, and finally scale it in such a way that the kinetic energy is unity [8]. We remove the forcing and the dissipative terms (which correspond to the linear terms) in the governing equations. These lead to Euler equations with no forcing and an initial field of zero mean velocity. We simulate this system, follow the change in kinetic energy and note the time it takes for the energy to increase by 1% (t1 ), 50% (t50 ) and diverge (tdiv ). Our results are summarized in Table 4 in comparison to [8]. It is evident that the presented method has very good energy conservation properties. We beleive this is due to the inherent satisfaction of the continuity equation, imposed regularity of the solution at r = 0 and the efficiency of spectral methods in conserving the energy. 5 Drag Reduction via Spanwise Oscillations Drag reduction due to spanwise oscillations is a well known phenomenon [2,18, 3,10,22,21], and numerical studies routinely report a drag reduction on the order of 40%. Spanwise oscillations in the case of pipe flow are essentially azimuthal oscillations (about the pipe axis). When the spanwise oscillations are imposed with the frequency Ω and the amplitude A, the azimuthal velocity at the pipe wall satisfies vosc (1, θ , z,t) = A sin(Ω t). In order to study the effect of the spanwise oscillations using the present approach, we set v = vosc − rAsin(Ω t), so that v still satisfies the homogenous boundary condition at the pipe wall, which is satisfied by the bases employed. With this change of variables (1) is transformed into (15) and the imposed oscillation acts as a forcing term in the new system;

∂t u + (u · ∇)u = −∇p + Re1ref ∆ u + F

(15)

∇·u = 0 u(1, θ , z,t) = 0

where F is the forcing due to the oscillation of the pipe wall,   (A sin (Ω t)∂θ u − 2A sin (Ω t))v + r(Asin(Ω t))2 F =  rΩ A cos (Ω t) + 2uA sin (Ω t) + A sin (Ω t)∂θ v  A sin (Ω t)∂θ w

In this section, the drag reducing effect of spanwise wall oscillations in turbulent regime is investigated. The investigation is carried out in two different settings. One simulation is run at Reτ = 180, with the flow driven by a constant

12

1

0.95

0.9



0.85

cf cf0

0.8

sing rea Inc

am

p lit

ude

0.75

0.7

0.65

0.6

0.55

0.5 0

10

5

15

20

t Fig. 6 Time evolution of skin friction coefficient scaled by uncontrolled (A = 0) case. For oscillation amplitudes, A = 2, 5, 10, 15, 20, and frequency Ω = 6π

pressure gradient. Use of this setting enables the investigation of the effects of wall oscillations on bulk flow. In a separate simulation, the mass flux is kept constant and the driving pressure gradient becomes time dependent. In this setting the Reynolds number is set to Re = 4900, as in the uncontrolled case. Keeping the mass flux constant enables the observation of changes in wall shear, and the pressure gradient needed to achieve constant flow rate. 5.1 Constant Pressure Gradient Case In this case, friction scaling is employed at Reτ = 180. The simulation is performed with the same parameters as the fixed wall case, with various values of the oscillation amplitude A and oscillation frequency Ω . The time evolution of skin friction coefficient scaled by the value for uncontrolled case (A = 0) is given in Figure 6. In this plot we present the results for a fixed oscillation frequency of Ω = 6π (frequency used in [3]), and vary the oscillation amplitude A. The skin friction coefficient, defined by, cf =

τw , 1 2 ρ 2 Ub

decreases from a value of 0.0094 to 0.0061 at A = 20 in the mean (see Figure 6). This amounts to a drag reduction of 35% consistent with the results previously published [3]. At higher amplitudes, drag reduction seems to level in our computational domain length of 10 radii as reported by [22] where a 20 radii length

13

25

20

15

wmean 10

5

0 0

20

40

60

80

100

120

140

160

180

r+ Fig. 7 Mean streamwise velocity profiles and bulk velocity for controlled (A 6= 0) and uncontrolled (A = 0) cases: ’’ No control, ’△’ A=20, Ω = 6π . Ucl /UB UB /uτ Ucl /uτ cf c f /c f 0 Recl Re Reτ

No control 1.32 14.58 19.27 0.00941 1.000 6864 5200 180

A=5 1.31 15.215 19.878 0.00864 0.918 7156 5477 180

A = 10 1.31 16.93 22.1 0.00698 0.742 7922 6094 180

A = 20 1.34 18.08 24.28 0.00611 0.649 8722 6508 180

Table 5 Comparison of turbulence statistics for controlled and uncontrolled flow, for fixed oscillation frequency Ω = 6π

is used in a constant mass flux simulation. In Figure 7, we present the mean velocity profile for controlled and uncontrolled cases. A summary of our results is presented in Table 5. We compute a bulk velocity Ub of 14.58 in the case of uncontrolled flow. When spanwise oscillations are imposed on the system however, bulk velocity increases. At an oscillation amplitude of 20, the bulk velocity increases to a value of 18.08, which represents a 24% increase, this ia again consistent with [3], where a 27% increase in bulk flow is reported.

5.2 Constant Mass Flux Case In addition to the constant pressure gradient case, a simulation in which the mass flux is kept constant, has been performed at Re = 4900. For this simulation an

14

Ucl /UB UB /uτ Ucl /uτ cf c f /c f 0 Recl Re Reτ

No control 1.32 14.58 19.27 0.00941 1.000 6864 5200 180

Ω = 3π 1.33 16.66 22.08 0.00724 0.769 7977 5998 180

Ω = 4π 1.32 17.25 22.75 0.00675 0.717 8197 6210 180

Ω = 5π 1.33 17.09 22.68 0.00685 0.728 8168 6151 180

Ω = 6π 1.31 16.93 22.10 0.00698 0.742 7922 6094 180

Table 6 Comparison of turbulence statistics for controlled and uncontrolled flow, for fixed oscillation amplitude A = 10

oscillation amplitude of A = 0.25 and an oscillation frequency of Ω = 0.35 are chosen based on the underlying scaling. These values approximately correspond to A = 7 and Ω = 10, when scaled in friction units as in [22]. The results are summarized in Table 7 in comparison to [22]. It is evident that, the pressure gradient required to keep the bulk velocity constant is decreased significantly in the case of wall oscillations. In this case, the order of drag reduction is about %26 which is in good agreement with [22] who employ similar simulation parameters. No control Ucl /UB UB /uτ Ucl /uτ cf c f /c f 0 G/GP G/Gnc Recl Re Reτ

1.31 14.08 18.45 0.01008 1.000 3.17 1 6419 4900 174

A = 0.25 Ω = 0.35 1.34 16.37 21.80 0.0075 0.746 2.29 0.722 6542 4900 149

Quadrio et al. [22] 1.34 16.60 22.18 0.00726 0.736 6556 4900 148

Table 7 Comparison of turbulence statistics for controlled and uncontrolled flow in the case of constant mass flux. GP denotes the mean pressure gradient required for corresponding laminar flow.

6 Conclusions and Discussion The main advantage of using solenoidal bases in incompressible fluid flow problems is that the solution is strictly solenoidal (i.e. divergence-free) throughout the solution domain. This eliminates the possible effects of the errors associated with poor resolution of the continuity equation. In addition, the elimination of the pressure variable due to the projection onto the dual solenoidal bases, simplifies the algorithm by removing the need for a pressure solver. In contrast to the popular solenoidal bases obtained from Karhunen-Loeve (KL) analysis (also known as Principal Orthogonal Decomposition), no separate simulation stage is required in the generation of the current bases. KL bases are known to be optimal in the

15

energy sense. On the other hand, they are no longer optimal when used for offreference parameter values, while the current bases are parameter-independent. It is also shown that, the method is effective in calculating the turbulence statistics at moderate Reynolds numbers, and has excellent energy conserving properties. The main disadvantage of the analytic solenoidal bases is the limitation to simple geometries in the current formulation. However, there are domain decomposition approaches (see [7]) as a possible remedy. Acknowledgements This study was supported in part by the Scientific and Technical Research ¨ ˙ITAK) through research project 109M435. Council of Turkey (TUB

A Appendix: Solenoidal Bases We list the forms of the solenoidal basis functions employed in this study below. In all the cases d P2m denotes the Legendre polynomial of order 2m, D+ = D + 1/r, and D = dr is replaced by a suitable Legendre differentiation submatrix after discretization based on the Gauss-Legendre grid in the radial direction [25], [26]. Case I, l6=0 n=0: Physical “1-basis”: Dual “1-basis”: ! ! 0 0 (1) (1) v˜ = H v = rH 0 0 H =P2m

H =(1 − r2 ) P2m Physical “2-basis”:

Dual “2-basis”:

v(1) =

−ilrG 0 D+ (rG)

!



v

(1)

−ilG 0 D+ G

=

!

G =(1 − r2 ) P2m

G =(1 − r2 )2 P2m Case II, l=0 n=0: Physical “1-basis”:

(1)

Dual “1-basis”: =

0 rH 0

!



v(2) =

!

Dual “2-basis”: 0 0 H

!

H =(1 − r2 ) P2m Case III, l6=0 n6=0:

=

0 H 0

H =P2m

H =(1 − r2 ) P2m Physical “2-basis”:

(1)

v˜ (1) =

0 0 rH

H =P2m

!

16 Physical “1-basis”:   −inr(α −1) G v(1) =  D(rα G)  0 G =(1 − r2 )2 P2m

α = min(|n|, 2 − mod(n, 2))

Physical “2-basis”: 

Dual “1-basis”: v˜

(1)

 −inr(α −1) G =  D(rα G)  0 

G =(1 − r2 ) P2m

α =mod(n, 2) + 1

Dual “2-basis”: 0



v(2) =  −ilr(α +1) H  inrα H H =(1 − r2 ) P2m

α = min(|n|, 2 − mod(n, 2))

 0 v˜ (1) =  −ilrα H  inr(α −1) H 

H =P2m

α =2 − mod(n, 2)

Hence the boundary conditions (7), the divergence free condition (8), and the regularity at the pipe center are all satisfied. Further, the integrands in the inner product integrals (9) are even.

17

References 1. Boyd, J. P., Chebyshev and Fourier Spectral Methods. Dover Publications, Inc., second edition, 2000. 2. Choi, K., Graham, M., Drag reduction of turbulent pipe flows by circular-wall oscillation. Physics of Fluids, 10(1):7, 1998. 3. Duggleby, A., Ball, K. S., and Paul, M. R., The effect of spanwise wall oscillation on turbulent pipe flow structures resulting in drag reduction. Physics of Fluids, 19(12):125107, 2007. 4. Duggleby, A., Ball, K. S., Paul, M. R., and Fischer, P., Dynamical eigenfunction decomposition of turbulent pipe flow. Journal of Turbulence, 8(772815469), 2007. 5. Eckhardt, B., Schneider, T. M., Hof, B., and Westerweel, J., Turbulence Transition in Pipe Flow. Annual Review of Fluid Mechanics, 39(1):447–468, 2007. 6. Eggels, J. G. M., Unger, F., Weiss, M. H., Westerweel, J., Adrian, R. J., Friedrich, R., and Nieuwstadt, F. T. M., Fully developed turbulent pipe flow: a comparison between direct numerical simulation and experiment. Journal of Fluid Mechanics, 268:175–210, 1994. 7. Pasquarelli, F., Domain decomposition for spectral approximation to Stokes equations via divergence free functions. Applied Numerical Mathematics, pages 493–514, 1991. 8. Fukagata, K., Highly Energy-Conservative Finite Difference Method for the Cylindrical Coordinate System. Journal of Computational Physics, 181(2):478–498, 2002. 9. Hesthaven, J., Gottlieb, S., and Gottlieb, D., Spectral Methods for Time-Dependent Problems. Cambridge, 2007. 10. Jung, W. J., Mangiavacchi, N., and Akhavan, F.L., Suppression of turbulence in wallbounded flows by high-frequency spanwise oscillations. Physics of Fluids A, 4(8):1605– 1608, 1992. 11. Kopriva, D. A., Implementing Spectral Methods for Partial Differential Equations. Springer, 2009. 12. Leonard, A., Wray A., A New Numerical Method for the Simulation of Three-Dimensional Flow in a Pipe, NASA Technical Memorandum 84267, 1982. 13. Meseguer, A., Trefethen, L. N., A spectral Petrov-Galerkin formulation for pipe flow I: Linear Stability and transient growth, Rep. 00/18, Numerical Analysis Group, Oxford University, Computing Lab., 2000. 14. Meseguer, A., Trefethen, L. N., A spectral Petrov-Galerkin formulation for pipe flow: II Nonlinear transitional stages, Rep. 01/19, Numerical Analysis Group, Oxford University, Computing Lab. 2001., 15. Meseguer, A., Trefethen, L. N., Linearized pipe flow to Reynolds number 10ˆ7. Journal of Computational Physics, 186(1):178–197, 2003. 16. Mhuiris, N. M. G., The construction and use of divergence free vector expansions for incompressible fluid flow calculations, ICASE Report No. 86-20, 1986. 17. Moser, R., Moin, P., and Leonard, A., A spectral numerical method for the Navier-Stokes equations with applications to Taylor-Couette flow. Journal of Computational Physics, 52(3):524–544, 1983. 18. Nikitin, N. V., On The Mechanism of Turbulence Suppression by Spanwise Surface Oscillations. Fluid Dynamics, (2):185–190, 2000. 19. Pasquarelli, F., Quarteroni, A., and Sacchi-Landriani, G., Spectral approximations of the stokes problem by divergence-free functions. J. Sci. Comput., 2(3):195–226, 1987. 20. Priymak. V. G,, Miyazaki, T., Accurate Navier-Stokes Investigation of Transitional and Turbulent Flows in a Circular Pipe. Journal of Computational Physics, 142(2):370–411, 1998. 21. Quadrio, M., Ricco, P., Critical assessment of turbulent drag reduction through spanwise wall oscillations. Journal of Fluid Mechanics, 521:251–271, 2004. 22. Quadrio, M., Sibilla, S., Numerical simulation of turbulent flow in a pipe oscillating around its axis. Journal of Fluid Mechanics, 424:217–241, 2000. 23. Reynolds, O., An Experimental Investigation of the Circumstances Which Determine Whether the Motion of Water Shall Be Direct or Sinuous, and of the Law of Resistance in Parallel Channels. Philosophical Transactions of the Royal Society of London, 174(1883):935– 982, 1883. 24. Schneider, T., Eckhardt, B., and Vollmer, J., Statistical analysis of coherent structures in transitional pipe flow. Physical Review E, 75(6), 2007.

18

25. Trefethen, L.N., Spectral Methods in Matlab. Society for Industrial and Applied Mathematics, Philadelphia, 2000. 26. Tu˘gluk, O., Tarman, H. I. Solenoidal bases for numerical studies of transition in pipe flow. Physica Scripta, T142:014009, 2010.

Direct Numerical Simulation of Pipe Flow Using a ...

These DNS are usually implemented via spectral (pseu- dospectral ... In this domain, the flow is naturally periodic along azimuthal (θ) direction and taken to be ...

157KB Sizes 0 Downloads 305 Views

Recommend Documents

DIRECT SIMULATION OF BLOOD FLOW IN A ...
the arterial wall or in order to slow vortices in the aneurysm and to favor coagulation of the sac. The stent is a complex multi-scale geometric structure that ...

Numerical Simulation of Heat Transfer and Fluid Flow ...
other types of fuel cells have to rely on a clean supply of hydro- gen for their operation. ... cathode and the anode is related to the Gibbs free energy change.

Simulation of blood flow in deformable vessels using ...
Jul 20, 2010 - models from medical imaging data have improved substantially in the last ..... real-time visualization of segmentation results. ...... 14th Annual Conference on Computer Graphics and Interactive Techniques, 1987; 163–169.

CFD Simulation of Heat Transfer in Turbulent Pipe Flow
radial variation of axial velocity, the turbulent kinetic energy, and the eddy diffusivity (compared with near- wall experimental data of Durst et al.5). The fourth.

Direct numerical simulations of particle transport in a ...
The discretization and the numerical solution procedure are briefly described in ..... If the fluid velocity u is known in advance, the analytic solution of Equation (6) ...

1630 Numerical Simulation of Pharmaceutical Powder Compaction ...
1630 Numerical Simulation of Pharmaceutical Powder Compaction using ANSYS - A Baroutaji.pdf. 1630 Numerical Simulation of Pharmaceutical Powder Compaction using ANSYS - A Baroutaji.pdf. Open. Extract. Open with. Sign In. Main menu. Displaying 1630 Nu

Numerical Simulation of Nonoptimal Dynamic ...
Computation of these models is critical to advance our ..... Let us now study the model specification of Example 2 in Kubler and Polemarchakis ..... sequentially incomplete markets, borrowing constraints, transactions costs, cash-in-advance.

Using the Direct Simulation Monte Carlo Approach for ...
The viability of using the Direct Simulation Monte Carlo (DSMC) approach to study the blast-impact ... by computing load definition for two model geometries - a box and an 'I' shaped beam. ... On the other hand, particle methods do not make the conti

Numerical simulation of saltwater upconing in a porous ...
Nov 9, 2001 - Grid convergence in space and time is investigated leading to ... transient, density-dependent flow field, and the experimental data are obtained ..... tured meshes is inferior to hexahedral meshes both with respect to memory.

Flow Simulation 2009 Tutorial
Analyze a Design Variant in the Flow Simulation Application . ...... name after the value and Flow Simulation will automatically convert the entered value to the SI ...

Flow Simulation 2009 Tutorial
Define the Engineering Goal . .... Define the Porous Medium - Unidirectional Type . ...... ffl We will use the Flow Simulation Analysis Tree to define our analysis ...

Numerical Aspects of Density Driven Flow in Porous ...
Density-driven flow in porous media is relevant in a number of frequently .... vanishes and a grid Peclet number Pg associated with the virtual velocity of. Pg =.

Numerical Aspects of Density Driven Flow in Porous ...
1 Interdisciplinary Center for Scientific Computing (IWR), Im Neuenheimer Feld 368, ... issues of the mathematical model and the design of fast solution methods ...

Conceptual and Numerical Flow Model of the Sines ...
the regional discharge area of the deep aquifer in the offshore. ... for the best possible reproduction of the aquifer equipotential surface, accommodating.

Numerical investigation of heat transfer and fluid flow ...
List of symbols cp. Specific heat, J/(kg K). D. Hydraulic diameter (=2H), mm f ... Local heat flux, W/m2 ..... gion, corresponding to local values at points F and G.

[B790.Ebook] Fee Download Partially Full Pipe Flow ...
Feb 18, 2012 - ... 2) By Harlan Bengtson theme around, why don't you read the book ... to go to the link that we supply and also ready to download and install.

Numerical modeling of unidirectional stratified flow with ...
fully developed exact solutions for unidirectional stratified flow. The evolutions of the interface in the ...... ification of solder material, Adv. Electron. Packaging 1.

Numerical simulation of three-dimensional saltwater ...
Dec 18, 2005 - subject to long-standing investigation and numerical analysis. ..... the software package d3f, a software toolbox designed for the simulation of.

Numerical Simulation of Fuel Injection for Application to ...
Simulate high speed reacting flow processes for application to Mach 10-12 .... 12. 14 x/d y/d. McDaniel&Graves. Rogers. Gruber et al. Musielak. Log. (Musielak) ...