Vous êtes sur la page 1sur 9

Imran Akhtar1

Interdisciplinary Center for Applied Mathematics, MC 0531, Virginia Tech, Blacksburg, VA 24061 e-mail: akhtar@vt.edu

A van der PolDufng Oscillator Model of Hydrodynamic Forces on Canonical Structures


Numerical simulations of the ow past elliptic cylinders with different eccentricities have been performed using a parallel incompressible computational uid-dynamics (CFD) solver. The pressure is integrated over the surface to compute the lift and drag forces on the cylinders. The numerical results of different cases are then used to develop reducedorder models for the lift and drag coefcients. The lift coefcient is modeled with a generalized van der PolDufng oscillator and the drag coefcient is expressed in terms of the lift coefcient. The parameters in the oscillator model are computed for each elliptic cylinder. The results of the model match the CFD results not only in the time domain but also in the spectral domain. DOI: 10.1115/1.3192127

Osama A. Marzouk Ali H. Nayfeh


University Distinguished Professor Department of Engineering Science and Mechanics, MC 0219, Virginia Tech, Blacksburg, VA 24061

Introduction

The interactive uid-structure phenomenon of ow separation and bluff body wakes has its fundamental signicance in ow physics and its practical importance in aerodynamic and hydrodynamic applications. When ow passes over a bluff body at low Reynolds numbers, ow separation may take place over substantial parts of its surface, but the ow around it remains steady. As the Reynolds number exceeds a critical value, instability in the separated shear layers develops, and nonlinear interaction of these layers with feedback from the wake leads to an organized and periodic motion of a regular array of concentrated vorticity, known as the von Krmn vortex street, in the wake. This vortex shedding exerts oscillatory forces on the body, which are often decomposed into drag and lift components along the freestream and cross-ow directions, respectively. If the body is capable of exing or moving rigidly, these forces can cause it to oscillate and the classical vortex-induced vibration VIV problem takes place. If the frequency of vortex shedding is close to a natural frequency of the body, the resulting resonance can generate large-amplitude oscillations, which may ultimately cause structural failure. Understanding this problem is of great interest in the design and maintenance of offshore structures, such as mooring cables, risers, and spars, and of high-aspect ratio structures subject to air streams, such as chimneys, high rise buildings, bridges, and cablesuspension systems. Since Roshko 1 measured the vortex-shedding period behind a bluff body, many researchers have investigated this phenomenon experimentally and numerically for a wide range of Reynolds numbers. The most frequently investigated bluff geometry is the circular cylinder. The ow behind a circular cylinder has become the canonical problem for studying such external separated ows 27 . Engineering applications, on the other hand, often involve ows over complex bodies, such as wings, submarines, missiles, and rotor blades, which can hardly be modeled as circular cylinders. Circular cylinders are extensively used in the study of bluff body uid dynamics due to their geometric simplicity and common use in engineering applications. Bishop and Hassan 2 were
1 Corresponding author. Contributed by the Design Engineering Division of ASME for publication in the JOURNAL OF COMPUTATIONAL AND NONLINEAR DYNAMICS. Manuscript received January 15, 2008; nal manuscript received February 12, 2009; published online August 25, 2009. Review conducted by Harry Dankowicz. Paper presented at the 46th AIAA Aerospace Sciences Meeting and Exhibit 2008, Paper No. AIAA-2008-679.

among the earliest to suggest using a self-excited oscillator to represent the forces over a cylinder due to vortex shedding. Hartlen and Currie 8 formulated a model for elastically restrained cylinders that are restricted to cross-ow motions. They used a Rayleigh oscillator to describe the lift force and coupled it to the cylinder motion by a linear velocity term. Currie and Turnball 9 proposed a similar model for the uctuations in the drag. Skop and Grifn 10 pointed out that the parameters in the model of Hartlen and Currie 8 lacked clear connection to physical parameters of the problem. They proposed a modied van der Pol oscillator to represent the lift coupled to a linear equation of motion for the structure. They also introduced the SkopGrifn parameter, which is now commonly used in studying VIV problems. Iwan and Blevins 11 considered the uid mechanics of the vortex street and developed a model in terms of a uid variable that captures the uid-dynamics effects of the problem. Landl 12 added a nonlinear aerodynamic damping term of fth order to the van der Pol oscillator in his two-equation model, suggesting that this enables better capturing of some physical characteristics. However, the model involves many constants to be determined. Evangelinos et al. 13 used direct numerical simulation DNS of the incompressible NavierStokes equations to solve, at a Reynolds number 1000, for the uid-structure problem of rigid and exible cylinders allowed to move freely in the cross-ow direction. The exible cylinder or beam was represented by a simplied linear model with a forcing term proportional to the lift coefcient and no structural damping. The governing equations were transformed into the Fourier domain and solved with a spectral elements method that employs a hybrid grid in the xy-plane and Fourier complex modes in the z-direction. They considered short and long cylinders with length-to-diameter ratios of 4 and 378, respectively. They reported that the often-used empirical formula proposed by Skop et al. 14 overpredicts the drag coefcient. Norberg 15 presented an overview of the uctuating pressure and lift acting on a circular cylinder, taking special consideration of the inuence of the Reynolds number and the relation between the uctuating lift and ow features e.g., laminar shedding, wake transition, and turbulent shedding in the near-wake region. He compiled the then available data from the diverse experimental and numerical approaches, including two-dimensional and threedimensional simulations, and reviewed the different measurement methods such as the force-element method, total force method, electromagnetic method, and the family of momentum pressure methods. These results show signicant changes in the uctuating pressure distribution over the cylinder for Reynolds numbers OCTOBER 2009, Vol. 4 / 041006-1

Journal of Computational and Nonlinear Dynamics Copyright 2009 by ASME

Downloaded 10 Nov 2011 to 128.173.39.66. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

= /2

U =

Ly

Lx

=0 =2

(a)

=3 /2
(b)

Fig. 1 a Geometry of an elliptic cylinder and b O-grid distributed among eight processors in the -direction

ranging from 47 to 200,000. A spanwise correlation length of about 30 cylinder diameters was observed for ows near the subcritical regime Re= 30,000. Williamson and Govardhan 16 gave a detailed review of the experimental and computational work on the VIV problem up to 2004. The primary concern was with the free and forced oscillations of elastically mounted rigid cylinders and bodies having two degrees of freedom and the dynamics of cantilevers, pivoted cylinders, cables, and tethered bodies. They underlined some of the fundamental questions in VIV and touched upon some debatable concepts, such as the added mass and the effect of combined mass-damping parameter on the peak-amplitude data in the Grifn plot. Gabbai and Benaroya 17 presented a comprehensive review of the studies experimental, semi-experimental, and numerical related to VIV problems. They covered a variety of issues related to the physics of the problem, such as the dynamics of a vibrating cylinder in a ow, the vortex-shedding modes, and the threedimensionality effects. They also reviewed different models e.g., wake models and force-decomposition models and approaches e.g., variational approach, vortex-in-cell VIC approach, DNS, and the nite-element method used to simulate the problem of VIV of both rigid and elastic cylinders. In the present paper, we present a numerical methodology capable of solving the ow past elliptic cylinders. We vary the cylinder eccentricity from = 0.5 to = 1.0 with an increment of 0.1 and validate/verify our simulation results by comparing them with other experimental and numerical results. We compute the lift and drag coefcients and analyze them using higher-order spectral moments to obtain the dominant frequencies and their relative phases. Using the spectral data, we identify the parameters in reduced-order models for the lift and drag. One of the potential applications would be to build a database of the model parameters for different Reynolds numbers and eccentricities and use it to predict the induced loads on cylinderlike and ellipselike structures.

1 p ui u jui = + + xi t xj

ui xj xj

where i , j = 1 , 2 , 3; the ui represent the Cartesian velocity components u , v , w ; p is the pressure; is the uid density; and is the uid kinematic viscosity. Equations 1 and 2 are nondimensionalized using the cylinder diameter D as the length scale and the freestream velocity U as the velocity scale. Thus, the Reynolds number is given by ReD = DU / . The other important geometric quantity, spanwise length of the body, is nondimensionalized using the cylinder diameter D and is denoted by Lz. We use a body conformal O type grid to simulate the ow over a body and employ curvilinear coordinates , , in an Eulerian reference frame; a planar view is shown in Fig. 1 a . Here, X and Y indicate the Cartesian coordinates and Lx and Ly are the major and minor axes, respectively. The thickness ratio of the ellipse is given by = Ly / Lx. This choice of coordinate system allows us to simulate the ow past any arbitrary closed shape, such as circular and elliptic cylinders and airfoils. Equations 1 and 2 are transformed into curvilinear coordinates in a strong conservative form as follows: Um
m

=0

J1ui Fim =0 + t m where the ux is dened as Fim = Umui + J1


m

xi

1 mn ui G ReD n

Numerical Methodology

The NavierStokes and continuity equations are the governing equations for the present problem. For incompressible ows, they can be represented as follows: ui =0 xi 041006-2 / Vol. 4, OCTOBER 2009 1

Here, J1 = det xi / j is the inverse of the Jacobian or the volume of the cell, Um = J1 m / x j u j is the volume ux contravariant velocity multiplied by J1 normal to the surface of constant m, and Gmn = J1 m / x j n / x j is the mesh skewness tensor. A non-staggered-grid layout is employed to solve the transformed NavierStokes equations. The Cartesian velocity components u , v , w and pressure p are dened at the center of the control volume in the computational space and the volume uxes U , V , W are dened at the midpoints of its corresponding faces. All of the spatial derivatives are approximated with second-order accurate central differences except for the convective terms. Using the same central differencing for the convection terms may lead to spurious oscillations in the coarser regions of the grid, thereby Transactions of the ASME

Downloaded 10 Nov 2011 to 128.173.39.66. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

leading to erroneous results. In the present formulation, we discretize the convective terms using a variation of quadratic upstream interpolation for convective kinematics QUICK ; that is, we calculate the face values of the velocity variables ui from the nodal values using a quadratic upwinding interpolation. The upwinding of QUICK is carried out by computing the positive and negative volume uxes Um + Um / 2 and Um Um / 2 , respectively, and using the generic stencil. A semi-implicit scheme is employed to advance the solution in time. The diagonal viscous terms are advanced implicitly using the second-order accurate CrankNicolson method, whereas all of the other terms are advanced using the second-order accurate AdamsBashforth method. The AdamsBashforth scheme was chosen because of its computational efciency when coupled with the fractional-step method. The discretized equations are Um
m

=0

J1

uik+1 uik 3 k 1 = Ci + DE uik Cik1 + DE uik1 + Ri pk+1 2 2 t + 1 DI uik+1 + uik 2 7

for a simple way to implement boundary conditions and keeps an equal load distribution for each processor. Moreover, the data points are exchanged only in one direction i.e., -direction , thereby reducing the interprocessor communication cost. For a three-dimensional simulation, the grid is divided in the - and -directions. Details of the solution algorithm and parallel implementation can be found in Refs. 20,21 . Because an application of suitable and well-posed boundary conditions is crucial for any simulation, we apply appropriate boundary conditions on different sections of the domain boundary. For the inow boundary condition, we use a Dirichlet boundary condition. For the outow boundary condition, we use a Neumann boundary condition. No-slip and no-penetration boundary conditions are applied on the cylinder surface. Moreover, periodic boundary conditions are applied in the - and -directions. The uid force on a cylinder is the manifestation of the pressure and shear stresses acting on its surface. The net force can be decomposed into two components: lift and drag forces. These forces are nondimensionalized with respect to the dynamic pressure. For the two-dimensional case, the coefcients of lift and drag can thus be written in terms of the dimensional pressure and shear stresses as follows: CL = 1 U 1 U2
2 0 2 2

p sin

where / m represent the discrete nite-difference operators in the computational space, the superscripts represent the time step, Ci represent the convective terms, Ri are the discrete operators for the pressure gradient terms, DI is the discrete operator representing the implicitly treated diagonal viscous terms, and DE is the discrete operator for the explicitly treated off-diagonal viscous terms. Mathematically, these terms are dened as follows: Ci =
m m

1 Re 1 Re

cos

12a

CD = where surface.
z

p cos +
0

sin

12b

is the spanwise vorticity component on the cylinder

U mu i

Validation and Verication

Ri =
m

J1

xi for m=n

DI =
m

Gmn
n

10

DE =
m

Gmn
n

for

11

It is important to note that, due to the orthogonality property for the specic case of a cylinder, the cross terms for the mesh skewness tensor Gmn are zero, that is, when m n. Therefore, the terms in Eq. 11 are identically zero and the stencil for the problem is a seven-point stencil. In the present formulation, we apply a fractional-step method 18 to advance the solution in time. The fractional-step method splits the momentum equation into a an advection-diffusion equationmomentum equation solved without the pressure term and b a pressure Poisson equationconstructed by implicit coupling between the continuity equation and the pressure in the momentum equation, thus satisfying the constraint of mass conservation. The governing equations are solved using a methodology similar to that employed by Zang et al. 19 . However, the algorithm is extended to parallel computing platforms and the 2D domain decomposition technique is employed to distribute the problem among different processors. To implement the algorithm on a distributed-memory, message-passing parallel computer, we use a two-dimensional domain decomposition technique such that each processor gets a slice of the grid, as shown in Fig. 1 b . In this gure, a two-dimensional view of 192 256 grid is shown, which is divided among eight processors such that each processor has a computational load of 192 32 grid points. This technique allows Journal of Computational and Nonlinear Dynamics

We validate the CFD results by comparing them with other experimental and numerical studies. We also perform grid and domain dependence study to verify the numerical results. We then perform numerical simulations of the ow past elliptic cylinders with thickness ratios varying from = 0.6 to = 1.0 with an incregrid ment of 0.1. The ow is simulated over a 192 256 with an outer domain size of 30D. In the present case, the grid is divided into eight domains/processors, such that each processor has a load of 192 32 grid points, as shown in Fig. 1 b . Using Eqs. 12a and 12b , we compute the lift and drag coefcients, respectively, for each cylinder and analyze them using higherorder spectral moments to obtain the dominant frequencies and their relative phases. We use these data to compute the parameters of the van der PolDufng oscillator model. 3.1 Validation. To validate the CFD results, we simulate the ow past elliptic cylinders. We compare the peak-to-valley lift p coefcient CL v, the mean drag coefcient CD, the Strouhal number St, and the mean surface pressure coefcient C P on the cylinder for the following two cases: Case I: = 0.5 elliptic cylinder Case II: = 1.0 circular cylinder In Case I, we simulate the ow past an elliptic cylinder with = 0.5 and a Reynolds number of 525 based on the projected length Ly of the elliptic cylinder. The numerical results are prep sented in Table 1. Our results for CL v, CD, and St are in good agreement with those of Mittal and Balachandran 22 . Also, we plot in Fig. 2 a the mean pressure coefcients C P over the surface of the elliptic cylinder. The pressure distribution is normalized such that C P = 1.0 at the stagnation point; that is, = 180. The C P distribution compares well with that of Mittal and Balachandran 22 . In Case II, we also simulate the ow past a circular cylinder at OCTOBER 2009, Vol. 4 / 041006-3

Downloaded 10 Nov 2011 to 128.173.39.66. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Table 1 Validation for the two cases Case I II Study Mittal and Balachandran 22 Present study Mittal and Balachandran 22 Present study
p CL v

Table 2 Grid and domain dependence studies St 0.2a 0.216 0.22 0.225 Grid size 192 288 216 256 384 256 Domain size 30D 30D 50D CD 1.415 1.429 1.381 St 0.2246 0.2246 0.2246

CD 0.78 0.766 1.44 1.415

1.21 1.05 2.42 2.366

Re= 1000.

Re= 525 on a 192 256 grid. In the simulation, the twodimensional domain has a diameter of 30D. For the specic case of a circular cylinder Lx = Ly = D. Similar to Case I, we compute the physical parameters associated with the ow and compare the results with those of Mittal and Balachandran 22 in Table 1. We also compare the C P distribution over the cylinder surface in Fig. 2 b and observe good agreement between the two numerical results. 3.2 Grid and Domain Dependence Studies. Grid and domain independence studies are critical for verifying the accuracy of the computational results. Therefore, we perform a grid/domain study for the ow past a circular cylinder at ReD = 525 and com pare some physical parameters, such as CD and St. We simulate the ow over a computational domain of 30D on eight processors with 192 256 and 288 384 grid points. The results from these simulations are given in Table 2. We observe that the percentage difference for CD and St is less than 5%, which clearly indicates that the ow in the vicinity of the structure is virtually grid independent. To establish domain independence, we increase the domain size by 60% to 50D and use 216 256 grid points. The slight increase in the grid points in the -direction, on the larger domain, is intentional to maintain comparable grid spacing in the two grids. The results from these simulations given in Table 2 indicate that the domain size of 30D is quite adequate for the current study.

parameters that varies across a critical value. Prototype selfexcited oscillators include the van der Pol and Rayleigh oscillators, which are often employed to model self-excitations in electrical and mechanical engineering applications. It is important to note that no autonomous single-degree-of-freedom equation, which does not include negative linear damping and positive nonlinear damping, can produce limit cycles. We showed 23 that the van der PolDufng oscillator model is applicable over a wide range of Reynolds number for the ow past a circular cylinder for the steady-state as well as transient lift coefcient and match the phase between dominant frequencies. In this paper, we extend the application of van der PolDufng oscillator to elliptical bodies with varying eccentricity. Nayfeh et al. 24 investigated two wake-oscillator models to represent the lift on a stationary circular cylinder. Analyzing the lift obtained with CFD using higher-order spectral moments, they found that the phase angle 13 between the two dominant lift components at f s and 3f s is around 90 deg. Based on this nding, they concluded that the van der Pol oscillator CL +
2 2 C L = C L C LC L

13

Reduced-Order Model

For the ow past a stationary bluff body, vortex shedding depends on many parameters, such as the Reynolds number, body geometry, and angle of attack. However, for a given set of parameters, the numerical and experimental results show that the long time history of the lift on the body is periodic and independent of the initial conditions. In other words, the ow is self-excited. Hence, when viewed as a dynamical system, the lift force can be represented by a self-excited oscillator. Using center manifold theory, one can reduce the dynamical system to an autonomous two-dimensional dynamical system or a single-degree-offreedom oscillator that exhibits a Hopf bifurcation as one of its

is more suitable for modeling the steady-state lift coefcient. The angular frequency in Eq. 13 is related but not equal to the angular shedding frequency s = 2 f s, the positive parameter represents negative damping, and the positive parameter represents positive nonlinear damping. The values of and are taken positive so that the linear damping is destabilizing while the nonlinear damping is stabilizing. As a consequence, small disturbances grow and large ones decay, both eventually approaching a stable limit cycle. The values of the parameters in Eq. 13 depend on the Reynolds number. Nayfeh et al. 25 found that the analytical model 24 reproduces the steady-state and the transient lift obtained with the CFD. They integrated Eq. 13 with different initial conditions and observed that the closer the initial conditions to the steady-state CFD values are, the better the matching between the results of the CFD and lift model is. Their results show that matching the transient lift also depends on the Reynolds number. At low Reynolds numbers, the transients due to the numerical simulation e.g., impulse or zero initial conditions decay much slower and the physical transient behavior of the dynamical system dominates later. However, at higher Reynolds numbers, the numerical transients decay

1 0.5

0.5
0

CP

CP

-0.5 -1 -1.5

-0.5

-1 0

90

180

270

360

90

180

270

360

(a)

(b)

Fig. 2 Mean surface CP distribution for a Case I and b Case II: present solid and Mittal and Balachandran 22 triangle

041006-4 / Vol. 4, OCTOBER 2009

Transactions of the ASME

Downloaded 10 Nov 2011 to 128.173.39.66. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

much faster and the van der Pol model predicts both of the transient and steady-state responses accurately. As an example, the nondimensional decay times of the numerical transients are approximately 50 and 16 for the Reynolds numbers 200 and 100,000, respectively. Nayfeh et al. 24 examined the spectra of the drag and lift obtained with CFD and found that the major peaks in the drag occur at 2f s and 4f s and that the phase between the drag peak at 2f s and the lift peak at f s is near 270 deg. Hence, they inferred that the periodic component of the drag is proportional to CLCL and proposed the following drag model: CD t = CD 2 a2 2 CL t CL t sa 1 14

where a2 is the amplitude of the drag component at 2f s and CD is the mean drag value. In the steady state, the mean component of the drag CD = CD ss is constant, while in the transient state, CD is a monotonically increasing function of time. The constant value CD ss is determined from the CFD steady-state time history of the drag and the value of a2 is determined from its spectral analysis. Qin 26 proposed that the quadratic term in the drag model 2 should be of the form CL instead of CLCL. He also found linear coherence between the drag and lift components at f s and 3f s and introduced a linear lift term in the drag model to account for it. Marzouk et al. 23 examined the phase angles 13 between the lift peaks at 3f s and f s more closely and found that it can vary by up to 25 deg around 90 deg, depending on the Reynolds number. To match this phase, they added a Dufng-type cubic term to the van der Pol oscillator model in Eq. 13 and obtained CL +
2 2 3 C L = C L C LC L C L

Fig. 3 Instantaneous spanwise vorticity contours

15

Moreover, they examined the phase 12 between the peak in the drag at 2f s and the peak in the lift at f s and found that it can vary up to 85 deg around 270 deg, again depending on the Reynolds number. To match this phase, they modied the drag model to CD t = CD + 2k1 a2 2 a2 2 CL CL + 2k2 2 C LC L a2 sa 1 1 16

In this paper, we investigate whether Eqs. 15 and 16 can model the lift and drag on elliptic cylinders. To this end, we express the solution of Eq. 15 using the method of harmonic balance as CL t = c1 cos
st

+ c2 sin

st

+ c3 cos 3 st + c4 sin 3 st 17

Substituting Eq. 17 into Eq. 15 , separating the terms, and multiplying the different sine and cosine functions yield the fourthorder linear system Ay = b 18
2 where y = 2 , , , , b = c1 , c2 , 9c3 , 9c4 s , and A = A ci , s . Next, we determine the ci by numerically simulating the ow past elliptic cylinders with different thickness ratios for ReLx = 525. We consider ve elliptic cylinders: = 0.6 Case 1 to = 1.0 Case 5 , with an increment of 0.1. All cases are simulated using eight processors and are run long enough to reach steady state. In Fig. 3, we plot the instantaneous spanwise vorticity for all cases. We observe a similar vortex-shedding pattern; however, the shedding frequency is different in each case. Moreover, the width of the wake increases with increasing , thereby increasing the drag. In other words, the size of the vortices being shed is of the order of Ly and the vorticity increases with . This is also evident from the fact that the projected area of the cylinders, as seen by the ow, increases with increasing . We then compute the time histories of the pressure distributions on the surfaces of the cylinders. Using Eqs. 12a and 12b , we

compute the lift and drag coefcients for each case. It follows from Figs. 4 a 4 e that the amplitudes of the uctuating forces increase as increases. Using spectral analysis, we compute the dominant frequencies and the phases 12 and 13. The results are given in Table 3. In Fig. 5, we plot the shedding frequency and the amplitudes a1, a2, and a3 of the dominant frequencies f s, 2f s, and 3f s, respectively. We note that odd harmonics appear in the lift spectrum, whereas even harmonics appear in the drag spectrum. In Fig. 5 a , we observe that the shedding frequency f s decreases with increasing , whereas the corresponding amplitude a1 increases as increases, as shown in Fig. 5 b . We observe approximately a vefold increase in a1 as increases from 0.6 to 1.0. Likewise, there is a similar trend in a2 and a3; however, the harmonics are much stronger in Case 5 than in Case 1 when compared with the respective a1. For example, in Case 1, a2 and a3 are 3.4% and 0.23% of a1, respectively, whereas in Case 5, they are 10.9% and 3%, respectively. Thus, the power distribution in the frequency spectrum is related to the projected area of the cylinder. We also note a steady increase in the mean drag, a direct consequence of the increase in the projected length of the cylinder, as shown in Fig. 5c. Next, we determine ci by matching Eq. 17 with the steadystate CFD solution, which is expressed as CL t = a1 cos
st

+ a3 cos 3 st +

13

19

Comparing Eq. 19 with Eq. 17 , we have c 1 = a 1, c2 = 0, c3 = a3 cos


13

c4 = sin

13

20

Substituting Eq. 20 into Eq. 18 and solving the resulting equation, we obtain y = 2 , , , and hence all of the parameters in Eq. 15 . Next, we integrate the identied van der PolDufng equation to predict the lift on the elliptic cylinders. In Fig. 6, we compare the time histories of the lift coefcient obtained using the model with those obtained with CFD for all of the cases. We observe good agreement between the CFD and model results. Next, we turn to the drag. The drag coefcient obtained with CFD can be expressed as CD t = CD + a2 cos 2 st + Moreover, substituting CL = a1 cos
st
12

21

into Eq. 16 yields

Journal of Computational and Nonlinear Dynamics

OCTOBER 2009, Vol. 4 / 041006-5

Downloaded 10 Nov 2011 to 128.173.39.66. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

0.5 0.4

0.6 0.4

0.3

CL , CD

CL , CD
305 310 315 320 325 330 335 340

0.2 0.1 0 0.1 0.2 300

0.2 0 0.2 0.4

(a)

T ime

(b)

300

305

310

315

T ime

320

325

330

335

340

1 0.8 0.6

CL , CD

0.2 0 0.2 0.4 0.6

CL , CD

0.4

0.5

0.5

(c)

300

305

310

315

T ime

320

325

330

335

340

(d)

300

305

310

315

T ime

320

325

330

335

340

1.5 1 0.5 0 0.5 1

CL , CD

(e)

300

305

310

315

T ime

320

325

330

335

340

Fig. 4 Time histories of the lift solid and drag dashed coefcients

CD t = CD

ss

+ a2 k1 cos 2 st k2 sin 2 st +

22

Comparing Eqs. 21 and 22 , we conclude that k1 = cos 12 and k2 = sin 12. Having determined k1 and k2, we compute the time histories of the drag coefcients using Eqs. 16 and 15 . The predicted time histories of the drag coefcient are compared with those obtained with CFD in Fig. 7. Again the agreement is excellent.

We list in Table 4 the model parameters for the ve cylinders. We observe that the angular frequency of the van der Pol Dufng oscillator model is close but not equal to the angular shedding frequency s. In Fig. 8 a , we plot the lift-model coefcients , , and as functions of . The parameters and are positive, indicating the presence of a limit cycle. It is clear from and that the inuence of the Dufng term, the values of

Table 3 Spectral analysis of the lift and drag coefcients Case 1 fs a1 a2 a3 CD


12 13 s

Case 2 0.273 0.425 0.0202 0.0028 0.66 25 deg 71 deg 1.715

Case 3 0.248 0.643 0.0437 0.0087 0.85 19 deg 86 deg 1.558

Case 4 0.231 0.891 0.0836 0.0182 1.08 21 deg 98 deg 1.451

Case 5 0.225 1.178 0.129 0.036 1.42 20 deg 97 deg 1.414

ss

0.318 0.238 0.0081 0.00055 0.51 30 deg 81 deg 1.998

041006-6 / Vol. 4, OCTOBER 2009

Transactions of the ASME

Downloaded 10 Nov 2011 to 128.173.39.66. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

0.4

1.2 1

0.3

a1, a2, a3
0.6 0.7 0.8 0.9 1

0.8 0.6 0.4 0.2 0

fs

0.2

0.1

(a)

0.6

0.7

0.8

(b)
180

0.9

1.5

90

12 , 13
0.6 0.7 0.8 0.9 1

CD

0.5

-90

-180

(c)
0.4

0.6

0.7

0.8

(d)
0.8

0.9

Fig. 5 Spectral analysis parameters from the CFD simulations


Model CFD Model CFD

0.2
L L

0.4

0.2

C
205 210 215 220

0.4

(a)

0.4 200

Time

(b)

0.8 200

205

Time

210

215

220

1 0.5
L

Model CFD

1.5 1 0.5
L

Model CFD

C
205 210 215 220

0 0.5 1

0 0.5 1

(c)

200

Time

(d)

1.5 200

205

Time

210

215

220

1.2

Model CFD

CL

0.2

0.8

(e)

1.8 200

205

Time

210

215

220

Fig. 6 Comparison between the time histories obtained by the lift model with those obtained with CFD

Journal of Computational and Nonlinear Dynamics

OCTOBER 2009, Vol. 4 / 041006-7

Downloaded 10 Nov 2011 to 128.173.39.66. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

0.528

Model CFD

0.69

Model CFD

0.52
D

0.512

C
0.504 200

0.67

0.65

(a)

205

Time

210

215

220

(b)

0.63 200

205

Time

210

215

220

0.94

Model CFD

1.25 1.2 1.15

Model CFD

0.9

CD

C
205 210 215 220

0.86

1.1 1.05

0.82 1 0.78 200 0.95 200 205 210 215 220

(c)

Time

(d)

Time

1.6

Model CFD

1.4

(e)

1.2 200

205

Time

210

215

220

Fig. 7 Comparison between the time histories obtained by the drag model with those obtained with CFD Table 4 Parameters of the reduced-order model Case 1 1.992 0.037 2.582 0.855 0.866 0.5 Case 2 1.671 0.084 1.831 1.146 0.906 0.423 Case 3 1.528 0.169 1.623 0.279 0.946 0.326 Case 4 1.484 0.236 1.198 0.137 0.934 0.358 Case 5 1.436 0.35 1.012 0.042 0.94 0.342

k1 k2

2.5 2

1 0.5

,,

1.5 1

k1 , k2
0.6 0.7 0.8 0.9 1

-0.5 0.5 0 -1

(a)

0.6

0.7

0.8

(b)

0.9

Fig. 8 Parameters of the van der PolDufng oscillator and drag models

041006-8 / Vol. 4, OCTOBER 2009

Transactions of the ASME

Downloaded 10 Nov 2011 to 128.173.39.66. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

although lesser in magnitude, cannot be neglected in the modeling. Likewise, we compute the coefcients k1 and k2 in the drag model, as shown in Fig. 8 b . The values of k1 and k2 suggest that 2 both of the terms CL and CLCL contribute signicantly to the modeling of the drag.

Conclusion and Future Work

The lift and drag coefcients for two-dimensional ows over elliptic cylinders were examined and modeled as a preliminary step to understanding the complex problem of vortex-induced vibrations. We extended the circular cylinder model for the lift and drag to elliptic cylinders. We performed numerical simulations over elliptic cylinders with varying eccentricities using a parallel CFD code. We studied ve cases of elliptic cylinders and calculated the hydrodynamic forces acting on them. We performed higher-order spectral analysis of the time histories of the lift and drag coefcients and computed the dominant frequencies and their relative phases. Using these data for different elliptic cylinders, we identied the coefcients in a van der PolDufng model for the lift and a drag model that depends quadratically on the lift. The model results predict well the CFD data. In the future, we will perform more simulations on elliptic cylinders with incoming ow at different angles of attack.

Acknowledgment
Numerical simulations were performed on Virginia Tech Advanced Research Computing-System X. The allocation grant and support provided by the staff is also gratefully acknowledged. Imran Akhtar would like to thank the Government of Pakistan for support during his graduate studies.

References
1 Roshko, A., 1955, On the Wake and Drag Of Bluff Bodies, J. Aero. Sci., 22 2 , pp. 124132. 2 Bishop, R., and Hassan, A., 1964, The Lift and Drag Forces on a Circular Cylinder in Flowing Fluid, Proc. R. Soc. London, Ser. A, 277, pp. 3250. 3 Karniadakis, G. E., and Triantafyllou, G. S., 1992, Three-Dimensional Dynamics and Transition to Turbulence in the Wake of Bluff Objects, J. Fluid Mech., 238, pp. 130. 4 Roshko, A., 1954, On the Development of Turbulent Wakes From Vortex Streets, NACA Report No. 1191. 5 Tomboulides, A. G., Israeli, M., and Karniadakis, G. E., 1989, Efcient Removal of Boundary Divergence Errors in Time-Splitting Methods, J. Sci. Comput., 4 3 , pp. 291308.

6 Williamson, C. H. K., 1996, Vortex Dynamics in the Cylinder Wake, Annu. Rev. Fluid Mech., 28, pp. 477539. 7 Wu, J., Sheridan, J., and Welsh, M. C., 1994, An Experimental Investigation of the Streamwise Vortices in the Wake of a Bluff Body, J. Fluids Struct., 8, pp. 621635. 8 Hartlen, R., and Currie, I., 1970, Lift-Oscillator Model of Vortex-Induced Vibration, J. Engrg. Mech. Div., 96, pp. 577591. 9 Currie, I., and Turnball, D., 1987, Streamwise Oscillations of Cylinders Near the Critical Reynolds Number, J. Fluids Struct., 1 2 , pp. 185196. 10 Skop, R., and Grifn, O., 1973, A Model for the Vortex-Excited Resonant Response of Bluff Cylinders, J. Sound Vib., 27 2 , pp. 225233. 11 Iwan, W., and Blevins, R., 1974, A Model for Vortex-Induced Oscillation of Structures, ASME J. Appl. Mech., 41 3 , pp. 581586. 12 Landl, R., 1975, A Mathematical Model for Vortex-Excited Vibration of Cable Suspensions, J. Sound Vib., 42 2 , pp. 219234. 13 Evangelinos, C., Lucor, D., and Karniadakis, G. E., 2000, DNS-Derived Force Distribution on Flexible Cylinders Subject to Vortex-Induced Vibrations, J. Fluids Struct., 14, pp. 429440. 14 Skop, R. A., and Balasubramanian, S., 1997, A New Twist on an Old Model for Vortex-Excited Vibrations, J. Fluids Struct., 11 4 , pp. 395412. 15 Norberg, C., 2003, Fluctuating Lift on a Cylinder: Review and New Measurements, J. Fluids Struct., 17, pp. 5796. 16 Williamson, C. H. K., and Govardhan, R., 2004, Vortex-Induced Vibrations, Annu. Rev. Fluid Mech., 36, pp. 413455. 17 Gabbai, R. D., and Benaroya, H., 2005, An Overview of Modeling and Experiments of Vortex-Induced Vibration of Circular Cylinder, J. Sound Vib., 282, pp. 575616. 18 Kim, J., and Moin, P., 1985, Application of a Fractional-Step Method to Incompressible Navier-Stokes, J. Comput. Phys., 59, pp. 308323. 19 Zang, Y., Street, R., and Koseff, J., 1994, A Non-Staggered Grid, Fractional Step Method for Time-Dependent Incompressible Navier-Stokes Equations in Curvilinear Coordinates, J. Comput. Phys., 114, pp. 1833. 20 Akhtar, I., 2008, Parallel Simulations, Reduced-Order Modeling, and Feedback Control of Vortex Shedding Using Fluidic Actuators, Ph.D. thesis, Virginia Tech, Blacksburg, VA. 21 Akhtar, I., Nayfeh, A. H., and Ribbens, C. J., 2009, On the Stability and Extension of Reduced-Order Galerkin Models in Incompressible Flows: A Numerical Study of Vortex Shedding, Theor. Comput. Fluid Dyn., 23 3 , pp. 213237. 22 Mittal, R., and Balachandar, S., 1995, Effect of Three-Dimensionality on the Lift and Drag of Nominally Two-Dimensional Cylinders, Phys. Fluids, 7 8 , pp. 18411865. 23 Marzouk, O. A., Nayfeh, A. H., Arafat, H. N., and Akhtar, I., 2007, Modeling Steadystate and Transient Forces on a Cylinder, J. Vib. Control, 13 7 , pp. 10651091. 24 Nayfeh, A. H., Owis, F., and Hajj, M. R., 2003, A Model for the Coupled Lift and Drag on a Circular Cylinder, ASME Paper No. DETC2003-VIB48455. 25 Nayfeh, A. H., Marzouk, O. A., Haider, H. N., and Akhtar, I., 2005, Modeling the Transient and Steady-State Flow Over a Stationary Cylinder, ASME Paper No. DETC2005-86376. 26 Qin, L., 2004, Development of Reduced-Order Models for Lift and Drag on Oscillating Cylinders With Higher-Order Spectral Moments, Ph.D. thesis, Virginia Tech, Blacksburg, VA.

Journal of Computational and Nonlinear Dynamics

OCTOBER 2009, Vol. 4 / 041006-9

Downloaded 10 Nov 2011 to 128.173.39.66. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Vous aimerez peut-être aussi