31.10.2014 Views

On Control Strategy for Fluid Flows using Model Reduction - Ibcast

On Control Strategy for Fluid Flows using Model Reduction - Ibcast

On Control Strategy for Fluid Flows using Model Reduction - Ibcast

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>On</strong> <strong>Control</strong> <strong>Strategy</strong> <strong>for</strong> <strong>Fluid</strong> <strong>Flows</strong> <strong>using</strong> <strong>Model</strong> <strong>Reduction</strong><br />

Imran Akhtar, Jeff Borggaard, and John Burns<br />

Interdisciplinary Center <strong>for</strong> Applied Mathematics, MC0531<br />

Virginia Tech, Blacksburg, VA 24061,Virginia, USA<br />

Email: jaburns@vt.edu , akhtar@vt.edu<br />

Abstract— The problem of active feedback control of fluid<br />

flows falls into a class of problems in the area of distributed<br />

parameter control, typically defined by partial-differential<br />

equations (PDEs). Physical processes modeled by PDEs are<br />

infinite dimensional systems are often simulated by numerical<br />

methods. However, <strong>for</strong> complex flows, the degrees of freedom<br />

may still be of the order of millions and are practical <strong>for</strong><br />

direct use in control design and optimization of fluid flow<br />

systems. Consequently, ”reduce-then-control” strategy is often<br />

employed <strong>for</strong> flow control of many engineering and industrial<br />

applications. In this study, we develop a linear quadratic<br />

regulator (LQR) control to suppress fluctuating <strong>for</strong>ces on a<br />

circular cylinder, through fluidic actuation, <strong>using</strong> a proper<br />

orthogonal decomposition (POD)-based low-dimensional model.<br />

We modify the model by applying suction on the cylinder<br />

surface and adding a control function in the velocity expansion.<br />

The nonlinear dynamical system thus developed is linearly<br />

unstable due to negative damping in the system. We linearize<br />

the system about the mean velocity and apply optimal control.<br />

We seek to minimize the fluctuating <strong>for</strong>ces on the cylinder<br />

<strong>using</strong> a reasonable amount of control ef<strong>for</strong>t. The novelty in this<br />

control strategy lies in feeding back only the dominant mode<br />

to suppress the fluctuating <strong>for</strong>ces. <strong>On</strong> the contrary, feedback of<br />

higher modes fails to control and destabilizes the system.<br />

I. INTRODUCTION<br />

<strong>Fluid</strong>-structure interaction often leads to undesirable hydrodynamic<br />

<strong>for</strong>ces ca<strong>using</strong> structural damage or even failure.<br />

<strong>Control</strong> of flow field in this perspective plays an important<br />

role and has been an active research area. Based on the control<br />

mechanism, flow control could be passive or active. Passive<br />

control devices, such as riblets, vortex generators, and<br />

boundary-layer trips, have been shown to be quite effective<br />

in delaying flow separation. Because passive devices can not<br />

adapt to flow changes, they might lose their efficiency during<br />

the process. <strong>On</strong> the other hand, active-control methods, such<br />

as acoustic excitation, vibrating ribbons or flaps, blowing and<br />

suction, couple the control input to the flow instabilities and<br />

can operate in a broad range of conditions.<br />

In order to develop a control strategy, it is important to<br />

understand the flow physics in the fluid-structure interaction.<br />

In general, researchers analyze a canonical flow initially <strong>for</strong><br />

physical insight in the flow field be<strong>for</strong>e modeling complex<br />

problems. Flow past a circular cylinder is one such example<br />

which has been studied extensively because of its canonical<br />

nature and being a typical unstable flow.<br />

When flow passes over a bluff body at low Reynolds<br />

numbers, flow separation may take place over substantial<br />

parts of its surfaces but the flow around it remains steady. As<br />

the Reynolds number exceeds a critical value, instability in<br />

the separated shear layers develops, and nonlinear interaction<br />

of these layers with feedback from the wake leads to an<br />

organized and periodic motion of a regular array of concentrated<br />

vorticity, known as the von Kármán vortex street in<br />

the wake, as shown in Figure 1. This vortex shedding exerts<br />

oscillatory <strong>for</strong>ces on the body, which are often decomposed<br />

into drag and lift components along the freestream and<br />

crossflow directions, respectively. If the body is capable<br />

of flexing or moving rigidly, these <strong>for</strong>ces can cause it to<br />

oscillate and the classical vortex-induced vibration (VIV)<br />

problem takes place. If the frequency of vortex shedding<br />

is close to a natural frequency of the body, the resulting<br />

resonance can generate large-amplitude oscillations, which<br />

may ultimately cause structural failure. Understanding this<br />

problem is of great interest in the design and maintenance<br />

of offshore structures, like mooring cables, risers, and spars,<br />

and of high-aspect ratio structures subject to air streams, like<br />

chimneys, high-rise buildings, bridges, and cable-suspension<br />

systems.<br />

Fig. 1.<br />

A typical von Kármán vortex street in the wake of a cylinder.<br />

Due to complexity in practical engineering problems of<br />

fluid flows, low-dimensional modeling is a widely used approach<br />

<strong>for</strong> engineers and researchers. Despite recent progress<br />

in control techniques, flow control remains a challenging<br />

task. Main limitation in designing a control rests in the<br />

infinite degree-of-freedom present in the governing equations<br />

in the <strong>for</strong>m of partial-differential equations (PDEs). Navier-<br />

Stokes equations provide one such example in which the<br />

Proceedings of International Bhurban Conference on Applied Sciences & Technology,<br />

Islamabad, Pakistan, 10 - 13 January, 2011<br />

77


presence of nonlinearity increases complexity of design and<br />

application of control. Low-dimensional modeling provides<br />

pratcial solutions <strong>for</strong> extremely challenging problems [1].<br />

<strong>Model</strong> reduction may be viewed as representing a physical<br />

phenomenon by a few number of equations or physically<br />

reducing the infinite or large dimensions of the problem. The<br />

<strong>for</strong>mer may be termed as phenomenological modeling while<br />

the latter is a more direct and conventional approach to model<br />

reduction.<br />

In phenomenological modeling, a flow-related phenomenon<br />

is modeled by a small degrees of freedom system.<br />

<strong>On</strong>e such example is vortex shedding in a bluff body wake.<br />

Several analytical approaches have been proposed to model<br />

vortex shedding over a circular cylinder. Bishop and Hassan<br />

[2] were among the earliest to suggest <strong>using</strong> a self-excited<br />

oscillator (Rayleigh or van der Pol oscillators) to represent<br />

the <strong>for</strong>ces over a cylinder due to vortex shedding. Several<br />

other analytical models were extended to elastic and moving<br />

cylinders [3], [4], [5], [6], [7]. Later, such models have been<br />

modified <strong>for</strong> various flow parameters [8](such as Reynolds<br />

numbers, <strong>for</strong>cing amplitudes etc.) or change in geometric<br />

parameters [9]. Other examples include modeling of heartbeat,<br />

current in RLC circuit, etc. <strong>using</strong> self-excited oscillator<br />

models.<br />

Many low-dimensional model techniques in fluid mechanics<br />

are derived from the proper orthogonal decomposition<br />

(POD)-Galerkin projection approach [10], [11]. The POD<br />

provides a tool to <strong>for</strong>mulate an optimal basis or minimum<br />

degrees of freedom (or modes) required to represent a<br />

dynamical system. The POD is also known as the Karhunen-<br />

Loeve expansion in statistics and principal component analysis<br />

or empirical orthogonal functions (EOF) in meteorology.<br />

The POD-Galerkin low-dimensional models are constructed<br />

in two steps:<br />

1) Computation of the POD basis functions from the data<br />

ensemble of the flow field.<br />

2) Galerkin projection of the Navier-Stokes equations<br />

onto a space spanned be a small number of POD basis<br />

functions.<br />

In the first step, the variables from the high-fidelity simulation<br />

(typically CFD) are transferred to a finite number<br />

of basis functions or modes, which are relatively small in<br />

number as compared to the degrees of freedom involved<br />

in the simulation. The second step involves a translation<br />

of the full-system dynamics to the implied dynamics of<br />

these modes. The resulting dynamical system consists of<br />

a set of ODEs in time in the modal amplitudes. Thus,<br />

low-dimensional models obtained from this procedure can<br />

be used to apply a control strategy. This application of<br />

control design at the analytical level will be beneficial <strong>for</strong><br />

the design of a controller in the real-time system. Some<br />

other applications of low-dimensional models include shape<br />

optimization, aeroelastic stability analysis, and understanding<br />

of the nonlinear dynamics of the system. A schematic of the<br />

hierarchy of flow models is represented in Figure 2 where<br />

the degrees of freedom are systematically reduced while<br />

preserving the physics and dominant features of the flow.<br />

A block diagram depicting the low-dimensional modeling proce-<br />

Fig. 2.<br />

dure.<br />

POD based low-dimensional models have been successfully<br />

implemented <strong>for</strong> various control strategies. In the current<br />

work, we would restrict ourselves only to the literature<br />

foc<strong>using</strong> on optimal control of vortex shedding past a circular<br />

cylinder. Graham et al. [12] developed a low-dimensional<br />

model <strong>for</strong> the flow past a circular cylinder at Re = 100 <strong>using</strong><br />

numerical simulations and the control action was achieved<br />

by cylinder rotation. They introduced two approaches to<br />

incorporate variable input control in the low-dimensional<br />

model. In the first approach, known as the “control function<br />

method,” a suitable control function is included in the velocity<br />

expansion to account <strong>for</strong> the inhomogeneous boundary<br />

conditions on the cylinder surface. The POD modes, used in<br />

the modified expansion, retain the homogeneous boundary<br />

conditions. In the second approach, the “penalty method,” the<br />

velocity expansion remains the same as <strong>for</strong> the unactuated<br />

flow and the essential boundary condition is en<strong>for</strong>ced in an<br />

integral “weak” fashion.<br />

Singh et al. [13] used these two approaches <strong>for</strong> controlling<br />

the cylinder wake. The models were based on POD of the<br />

two-dimensional incompressible, unsteady wake flow behind<br />

a circular cylinder at Re = 100. Linear optimal control theory<br />

was used to develop the feedback gains. Their simulations<br />

of the modes showed that, in a closed-loop system, asymptotic<br />

regulation of the amplitudes to the desired equilibrium<br />

state could be accomplished by appropriately rotating the<br />

cylinder. Also, they found that their linear controller was<br />

capable of suppressing large unsteady perturbations, despite<br />

the presence of nonlinearity in the flow dynamics of the two<br />

models.<br />

Bergmann et al. [14] employed optimal control approach<br />

Proceedings of International Bhurban Conference on Applied Sciences & Technology,<br />

Islamabad, Pakistan, 10 - 13 January, 2011<br />

78


to reduce the drag on the cylinder at Reynolds number of<br />

200 <strong>using</strong> cylinder rotation. They used the time angular<br />

velocity of the rotating cylinder as the cost function. They<br />

developed POD-based reduced order model and introduced<br />

time-dependent eddy-viscosity estimated <strong>for</strong> each POD mode<br />

as the solution of an auxiliary optimization problem. They<br />

used Lagrange multipliers to en<strong>for</strong>ce constraints to solve<br />

optimization problem and achieved 25% of relative drag<br />

reduction.<br />

Siegel et al. [15] numerically studied the effect of feedback<br />

flow control on the cylinder wake at Re = 100. The controller<br />

applies linear proportional and differential feedback to<br />

the estimate of the first POD mode. Actuation is implemented<br />

as the displacement of the cylinder normal to the flow<br />

while the sensors were placed in the wake. They observed a<br />

reduction of 15% and 90% of drag and lift, respectively.<br />

Bergmann and Cordier [16] used optimal control theory<br />

to minimize the mean drag <strong>for</strong> a circular cylinder with<br />

amplitude and frequency as the control parameters <strong>using</strong><br />

cylinder rotation. They employed trust-region POD approach<br />

in the wake which converged to the minimum predicted by<br />

an open-loop control approach and leads to a relative mean<br />

drag reduction of 30%.<br />

In this study, we follow an optimal control strategy to<br />

minimize the fluctuating <strong>for</strong>ces. Most of the studies mentioned<br />

have used cylinder rotation as the means of actuation.<br />

In this study, we use suction jet as the actuator. We also<br />

define the objective function based on the lift coefficient. We<br />

develop a POD-based low-dimensional model <strong>using</strong> a control<br />

function approach based on suction on the cylinder surface.<br />

We feedback only the dominant (first) state of the system and<br />

compute appropriate gains. Successful application of optimal<br />

control on the low-dimensional model suggests that the<br />

fluctuating <strong>for</strong>ces can be completely suppressed <strong>using</strong> suction<br />

actuators. The manuscript is organized as follows; Section<br />

II presents the numerical methodology used to simulate the<br />

flow past a circular cylinder. In Section III, we discuss the<br />

methodology to compute POD basis functions and compute<br />

them <strong>for</strong> the two-dimensional problem. In Section IV, we<br />

develop a reduced-order model <strong>for</strong> the unactuated flow field.<br />

Later, we discuss control function method and develop a<br />

reduced-order model <strong>for</strong> the actuated flow field. In Section<br />

V, we design an LQR control based on the linearized model<br />

and numerically integrate the system to show suppression of<br />

system response.<br />

II. NUMERICAL METHODOLOGY<br />

where the flux is defined as<br />

F im = U m u i + J −1 ∂ξ m<br />

∂x i<br />

p − 1 Re Gmn ∂u i<br />

∂ξ n<br />

. (3)<br />

−1 ∂ξm<br />

∂x j<br />

∂ξ n<br />

∂x j<br />

Here Re = U∞D<br />

ν<br />

; J −1 = det ( ∂x i<br />

)<br />

∂ξ j<br />

is the inverse of the<br />

−1 ∂ξm<br />

Jacobian or the volume of the cell; U m = J<br />

∂x j<br />

u j is the<br />

volume flux (contravariant velocity multiplied by J −1 ) normal<br />

to the surface of constant ξ m ; and G mn = J<br />

is the “mesh skewness tensor.”<br />

A semi-implicit scheme is employed to advance the solution<br />

in time. The diagonal viscous terms are advanced<br />

implicitly <strong>using</strong> the second-order accurate Crank-Nicolson<br />

method, whereas all of the other terms are advanced <strong>using</strong><br />

the second-order accurate Adams-Bash<strong>for</strong>th method. The<br />

Adams-Bash<strong>for</strong>th scheme was chosen because of its computational<br />

efficiency when coupled with the fractional step<br />

method. Details of the numerical algorithm, validation and<br />

verification can be found in Ref. [17], [9], [18]. <strong>Fluid</strong>ic<br />

actuation (suction, blowing, or synthetic jets) on the cylinder<br />

surface can be used as a controlling mechanism <strong>for</strong> the flow<br />

field. The actuator is modeled by imposing a fixed or timevarying<br />

velocity normal to the cylinder surface [19].<br />

The fluid <strong>for</strong>ce on a cylinder is the manifestation of the<br />

pressure and shear stresses acting on its surface. The net<br />

<strong>for</strong>ce can be decomposed into two components: lift and drag<br />

<strong>for</strong>ces. These <strong>for</strong>ces are nondimensionalized with respect to<br />

the dynamic pressure. The coefficients of lift and drag can<br />

thus be written in terms of the dimensional pressure and<br />

shear stresses as follows:<br />

C L = − 1<br />

ρU 2 ∞<br />

C D = − 1<br />

ρU 2 ∞<br />

∫2π<br />

(<br />

p sin θ − 1 )<br />

Re ω z cosθ dθ<br />

0<br />

∫2π<br />

(<br />

p cosθ + 1 )<br />

Re ω z sin θ dθ<br />

0<br />

(4a)<br />

(4b)<br />

where ω z is the spanwise vorticity component on the<br />

cylinder surface.<br />

III. POD BASIS FUNCTIONS<br />

A parallel CFD solver is used to simulate the flow past a<br />

circular cylinder [17], [18]. The governing equations are the<br />

Navier-Stokes equations in curvilinear coordinates (ξ m ) and<br />

can be written in strong conservation <strong>for</strong>m as follows:<br />

∂(J −1 u i )<br />

∂t<br />

∂U m<br />

∂ξ m<br />

= 0, (1)<br />

+ ∂F im<br />

∂ξ m<br />

= 0, (2)<br />

In this section, we provide the procedure <strong>for</strong> computing<br />

the POD basis functions <strong>for</strong> two-dimensional flow fields. The<br />

flow field data (u, v) is generated from an experiment or a<br />

numerical simulation and is assembled in a matrix W 2N×S ,<br />

as shown in equation (5); each column represents one time<br />

instant or a snapshot and S is the total number of snapshots<br />

<strong>for</strong> N grid points in the domain. The vorticity field can also<br />

be used <strong>for</strong> POD, however, in the case of the velocity field,<br />

the singular values are a direct measure of the kinetic energy<br />

Proceedings of International Bhurban Conference on Applied Sciences & Technology,<br />

Islamabad, Pakistan, 10 - 13 January, 2011<br />

79


in each mode.<br />

⎡<br />

W =<br />

⎢<br />

⎣<br />

u (1)<br />

1 u (2)<br />

1 . . . u (S)<br />

1<br />

. . .<br />

u (1)<br />

N<br />

u (2)<br />

N<br />

. . . u (S)<br />

N<br />

v (1)<br />

1 v (2)<br />

1 . . . v (S)<br />

1<br />

. . .<br />

. . .<br />

v (1)<br />

N<br />

v (2)<br />

N<br />

. . . v (S)<br />

N<br />

Mathematically, we compute Φ <strong>for</strong> which the following<br />

quantity is maximum:<br />

〈|u, Φ| 2〉<br />

⎤<br />

⎥<br />

⎦<br />

(5)<br />

‖Φ‖ 2 , (6)<br />

where Φ are the POD basis functions and 〈.〉 denotes the<br />

ensemble average. Applying variational calculus, one can<br />

show that equation (6) is equivalent to a Fredholm integral<br />

eigenvalue problem represented as<br />

In the classical POD or direct method, originally introduced<br />

by Lumley [20], a two-point spatial-correlation tensor<br />

is <strong>for</strong>med and the eigenfunctions are the POD modes. In<br />

this approach, the average operator is estimated in time. <strong>On</strong><br />

the other hand, if the average operator is evaluated as a<br />

space average over the domain of interest, the method is<br />

known as the method of snapshot [21]. In this approach, we<br />

<strong>for</strong>mulate a temporal-correlation function from the snapshots<br />

and trans<strong>for</strong>m it into an eigenvalue problem.<br />

In this study, we compute the singular value decomposition<br />

(SVD) of the data ensemble W to obtain the POD modes,<br />

i.e. W = UΣV T , where U represents the POD basis. This<br />

method has a limited application, especially when the grid<br />

size (N) is large. [22] proposed an algorithm, inspired<br />

from domain decomposition ideas, <strong>for</strong> a scalable parallel<br />

efficient computation of the POD basis vector with low<br />

communication overhead. An SVD of a (spatial) subdomain<br />

time history is calculated locally on each processor followed<br />

by the exchange of a small number of dominant V p (right singular<br />

vector on the p th processor) with other processors. An<br />

iterative application of this step showed promising results <strong>for</strong><br />

complex fluid flows, gravity currents, and two-dimensional<br />

flow past a square cylinder. Thus,<br />

⎡<br />

U =<br />

⎢<br />

⎣<br />

φ (u)<br />

1,1 φ (u)<br />

2,1 . . . φ (u)<br />

S,1<br />

⎤<br />

. . .<br />

φ (u)<br />

1,N φ(u) 2,N . . . φ(u)<br />

S,N<br />

φ (v)<br />

1,1 φ (v)<br />

2,1 . . . φ (v)<br />

S,1<br />

⎥<br />

. . . ⎦<br />

φ (v)<br />

1,N φ(v) 2,N . . . φ(v) S,N<br />

and Σ = diag[σ 1 σ 2 . . . σ S ] T are the singular values.<br />

An important characteristic of these modes is orthogonality;<br />

that is, Φ i .Φ j = δ ij , where δ ij is the kronecker delta. The<br />

optimality of the POD modes lies in capturing the greatest<br />

(7)<br />

possible fraction of the total kinetic energy <strong>for</strong> a projection<br />

onto the given set of modes. In equation (7), Σ contains the<br />

singular values of W. The corresponding singular values σ i<br />

are real and positive arranged in Σ in descending order. The<br />

σ i are related to the eigenvalues λ i obtained in the method<br />

of snapshots by λ i = σi 2 . They are used <strong>for</strong> ordering the<br />

POD basis functions <strong>for</strong> the dynamical system and represent<br />

energy contained in each basis function.<br />

Among other factors, both of the number of snapshots and<br />

the number of POD modes used in the simulation affect the<br />

accuracy of the reduced-order model. Increasing the number<br />

of modes increases the accuracy of the simulation, but it<br />

also increases the dimension of the solution space spanned<br />

by the modes. Similarly, increasing the number of snapshots<br />

increases the flow dynamics in<strong>for</strong>mation, however, it limits<br />

the computational power and available computer memory.<br />

IV. LOW-DIMENSIONAL MODELING<br />

In the current study, we per<strong>for</strong>m two-dimensional simulations<br />

of the flow past a cylinder at Re = 200 on a 192 ×256<br />

grid in the radial and circumferential directions, respectively,<br />

over the domain size of 50D. We record 40 snapshots of the<br />

flow field over a shedding cycle ans assemble the data in<br />

W matrix. We subtract the mean component and per<strong>for</strong>m<br />

SVD of the matrix to compute the POD eigenfunctions. In<br />

Figure 3, we plot these eigenvalues where i th eigenvalue is<br />

normalized as λ i / ∑ S<br />

j=1 λ j. First twelve POD modes contain<br />

99% of the system’s energy and are sufficient <strong>for</strong> lowdimensional<br />

modeling <strong>for</strong> this configuration. We also plot<br />

the first four POD modes of the streamwise (φ u ) and normal<br />

(φ v ) velocity components in Figures 4 and 5, respectively.<br />

λ<br />

10 0<br />

10 -1<br />

10 -2<br />

10 -3<br />

10 -4<br />

10 -5<br />

10 -6<br />

10 -7<br />

10 -8<br />

Mode<br />

Fig. 3. Normalized eigenvalues.<br />

5 10 15 20<br />

For the flow past a circular a circular cylinder, [23]<br />

observed that 20 snapshots are sufficient <strong>for</strong> the construction<br />

of the first eight eigenfunctions. Numerical studies [24]<br />

suggest that the first M POD modes, where M is even, resolve<br />

the first M/2 temporal harmonics and require 2M number<br />

of snapshots <strong>for</strong> convergence.<br />

Proceedings of International Bhurban Conference on Applied Sciences & Technology,<br />

Islamabad, Pakistan, 10 - 13 January, 2011<br />

80


Fig. 4.<br />

Fig. 5.<br />

(a) Mode 1 (b) Mode 2<br />

(c) Mode 3 (d) Mode 4<br />

The streamwise velocity modes (φ u i , i = 1, 2, 3,4) at Re=200.<br />

(a) Mode 1 (b) Mode 2<br />

(c) Mode 3 (d) Mode 4<br />

The normal velocity modes (φ v i , i = 1, 2,3, 4) at Re=200.<br />

A. Low-dimensional <strong>Model</strong> Without <strong>Control</strong><br />

In the unactuated low-dimensional model, the velocity<br />

field is expanded as<br />

u(x, t) ≈ ū(x) +<br />

M∑<br />

q i (t)Φ i (x), (8)<br />

i=1<br />

where M is the number of POD modes used in the projection.<br />

We substitute equation (8) into equation (2), project this<br />

equation onto the Φ k , and obtain<br />

q˙<br />

k (t) = A k +<br />

where<br />

M∑<br />

B km q m (t) +<br />

m=1<br />

M∑<br />

m=1 n=1<br />

A k = 1 Re (Φ k, ∇ 2 ū) − (Φ k ,ū.∇ū),<br />

M∑<br />

C kmn q n (t)q m (t), (9)<br />

B km = −(Φ k ,ū.∇Φ m ) − (Φ k , Φ m .∇ū) + 1 Re (Φ k, ∇ 2 Φ m ),<br />

C kmn = −(Φ k , Φ m .∇Φ n ),<br />

and (a, b) = ∫ Ω<br />

a · b dΩ represents the inner product<br />

between a and b. We note that <strong>using</strong> Green’s theorem<br />

and the divergence-free property, the pressure term drops<br />

out from equation (9) <strong>for</strong> the case of p=0 on the outflow<br />

boundary [25]. The POD basis functions are identically<br />

zero on the inflow boundary because the average flow is<br />

subtracted from the total flow. However, in case of Neumann<br />

boundary conditions on outflow boundary, the contribution<br />

of the pressure term is not exactly zero <strong>for</strong> the cylinder<br />

wake. The outer domain is intentionally kept at 25D from<br />

the cylinder to minimize the pressure effects. Hence, the<br />

pressure is neglected on the outflow boundary so the pressure<br />

term vanishes in the reduced-order model [24]. Details of the<br />

projection and validation of the low-dimensional model can<br />

be found in [17], [18].<br />

[18] developed a low-dimensional model of the pressure<br />

field by <strong>using</strong> the pressure-Poisson equation. Similar to the<br />

velocity field, the pressure field data is recorded during<br />

the CFD simulation. For incompressible flows, the pressure-<br />

Poisson equation<br />

∂ 2 p<br />

∂x 2 i<br />

= − ∂u i<br />

∂x j<br />

∂u j<br />

∂x i<br />

, (10)<br />

where i and j refer to the Cartesian components of the vector,<br />

is the governing equation <strong>for</strong> the pressure. The pressure field<br />

is then expanded as a sum of a mean component ( ¯P ) and a<br />

fluctuations component (p ′ ). The average pressure ¯P = 〈p〉<br />

is subtracted from the snapshot data and the pressure POD<br />

modes (Ψ i ) are computed. The pressure is then expanded in<br />

terms of the Ψ i (x) as follows:<br />

p(x, t) ≈ ¯P(x) +<br />

M∑<br />

a m (t)Ψ m (x). (11)<br />

m=1<br />

It is important to note that, in the Galerkin expansion in<br />

equation (11), the temporal coefficients a i (t) are different<br />

from the q i (t).<br />

Substituting equations (8) and (11) into equation (10) and<br />

projecting it onto the pressure POD modes yields the lowdimensional<br />

model <strong>for</strong> the pressure field<br />

D km a k (t) = E k +<br />

where<br />

M∑<br />

F km q m (t) +<br />

m=1<br />

D km = (Ψ k , ∇ 2 Ψ m )<br />

M∑<br />

m=1 n=1<br />

E k = −(Ψ k , ∇ 2 ¯P) − (Ψk , ∇ū : ∇ū)<br />

F km = −(Ψ k , ∇ū : ∇Φ m ) − (Ψ k , ∇Φ m : ∇ū)<br />

G kmn = −(Ψ k , ∇Φ m : ∇Φ n ).<br />

M∑<br />

G kmn q n (t)q m (t), (12)<br />

Equations (12) constitute a set of algebraic equations<br />

quadratic in terms of the q i . The pressure thus obtained is<br />

integrated over the cylinder surface to compute the lift and<br />

drag <strong>for</strong>ces. From equation (12), we clearly observe that<br />

where G is a nonlinear function of q.<br />

Proceedings of International Bhurban Conference on Applied Sciences & Technology,<br />

Islamabad, Pakistan, 10 - 13 January, 2011<br />

a = G(q) (13)<br />

81


In Figure 6, we plot the lift coefficient (C L ) and the<br />

first state (q 1 ) of the system. We observe that the two time<br />

histories have different magnitudes but approximately the<br />

same frequency. If we drive the response of q 1 to zero, it<br />

would suppress the lift <strong>for</strong>ce as well. Thus, we develop a<br />

modified low-dimensional model which includes the control<br />

mechanism and apply optimal control to drive the system<br />

response to zero.<br />

C L<br />

, q 1<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

100 110 120 130 140<br />

Time<br />

Fig. 6. Time histories of the lift coefficient (C L : solid) and first mode<br />

q 1 : dashed.<br />

B. Low-dimensional <strong>Model</strong> With <strong>Control</strong><br />

In the current study, we use a pair of suction actuators on<br />

the cylinder surface as a control mechanism. The location and<br />

suction velocity of the actuators is motivated by the study<br />

of [17] where they used pressure POD mode distribution on<br />

the cylinder surface to optimally locate the actuators. The<br />

location is approximately ±75 ◦ from the base point of the<br />

cylinder.<br />

We use a control function approach to develop a lowdimensional<br />

model incorporating the control . The velocity<br />

field is expanded as<br />

u(x, t) ≈ ū(x) +<br />

M∑ ∑M c<br />

q i (t)Φ i (x) + γ i (t)Γ i (x), (14)<br />

i=1<br />

i=1<br />

where M c is the total number of control modes, the Γ i (x) are<br />

a suitable divergence-free control functions that satisfy the<br />

inhomogeneous boundary condition due to fluidic actuators<br />

(suction), and γ(t) is the variable control input. Details of<br />

the derivation and construction of the control mode can be<br />

found in Ref. [17], [26], [19].<br />

We substitute equation (14) into the Navier-Stokes equations,<br />

project these equations onto the Φ k , and obtain<br />

˙q k (t) = A k +<br />

where<br />

+<br />

∑M c<br />

m=1<br />

M c<br />

M∑<br />

B km q m (t) +<br />

m=1<br />

H km ˙γ m (t) +<br />

M∑<br />

m=1 n=1<br />

M∑<br />

C kmn q n (t)q m (t)<br />

M∑ ∑M c<br />

J kmn q m (t)γ n (t)<br />

m=1 n=1<br />

M c M c<br />

∑ ∑ ∑<br />

+ K km γ m (t) + L kmn γ m (t)γ n (t), (15)<br />

m=1<br />

H km = −(Φ k , Γ m ),<br />

m=1 n=1<br />

J kmn = −(Φ k , Γ n · ∇Φ m ) − (Φ k , Φ m · ∇Γ n ),<br />

K km = −(Φ k ,ū · ∇Γ m ) − (Φ k , Γ m · ∇ū) + 1<br />

Re D<br />

(Φ k , ∇ 2 Γ m ),<br />

L kmn = −(Φ k , Γ m · ∇Γ n ).<br />

V. APPLICATION OF LQR CONTROL<br />

The low-dimensional model, developed in equation (15),<br />

is a 12-dimensional system of nonlinear ODEs. We simplify<br />

the system to apply optimal control technique in an ef<strong>for</strong>t to<br />

reduce the fluctuating <strong>for</strong>ces on the structure. We linearize<br />

the system about the mean velocity component such that<br />

q ∗ = 0 is the equilibrium point. According to the Hartman-<br />

Grobman theorem [27], (a) the fixed point, say q = q 0 ,<br />

of the nonlinear system of the <strong>for</strong>m in equation (9) is<br />

stable when the fixed point q ∗ = 0 of the linear system<br />

is asymptotically stable; and the fixed point, q = q 0 , of the<br />

nonlinear system of the <strong>for</strong>m in equation (9) is unstable when<br />

the fixed point q ∗ = 0 of the linear system is unstable. In a<br />

topological setting, the Hartman-Grobman theorem implies<br />

that the trajectories in the vicinity of a hyperbolic fixed point<br />

q = q 0 are qualitatively similar to those in the vicinity of the<br />

hyperbolic fixed point q ∗ = 0 of the linearized system. In<br />

other words, the local nonlinear dynamics near q = q 0 is<br />

qualitatively similar to the linear dynamics near q ∗ = 0, and<br />

the qualitative change in the local nonlinear dynamics can be<br />

detected by examining the associated linear dynamics. Thus,<br />

we linearize the nonlinear system and apply optimal control<br />

<strong>for</strong>mulation to make the system stable.<br />

This approach reduces the nonlinear system to a lineartime-invariant<br />

(LTI) system and allows us to effectively apply<br />

linear control theory. Thus the augmented LTI system can<br />

be written in the standard <strong>for</strong>m of state-space equations as<br />

follows: [ ] [ ] [ ] [ ]<br />

B K q H<br />

= ˙q˙γ 0 T + γ<br />

0 γ 1 c (16)<br />

Here, γ c = ˙γ is a new control input to the system. We<br />

design an optimal controller based on the linearized system<br />

and make it positively damped such that the response decays<br />

to zero.<br />

A. <strong>Control</strong> Objective<br />

As discussed earlier, vortex shedding leads to fluctuating<br />

<strong>for</strong>ces on the structure and may lead to VIV <strong>for</strong> elastic<br />

Proceedings of International Bhurban Conference on Applied Sciences & Technology,<br />

Islamabad, Pakistan, 10 - 13 January, 2011<br />

82


structures. Thus, we want to suppress vortex shedding and<br />

reduce the fluctuating <strong>for</strong>ces acting on the structure <strong>using</strong> a<br />

reasonable amount of control ef<strong>for</strong>t. From the flow control<br />

point of view, the optimal control of the Navier-Stokes equations<br />

through boundary control is determined to minimize the<br />

lift <strong>for</strong>ces on the cylinder. Mathematically, we seek<br />

min<br />

∫ ∞<br />

0<br />

CL 2 (t) + Rγ2 c (t) dt<br />

where R represents a control cost and the dynamics are<br />

described by the low-dimensional model derived in equation<br />

(15). We now use the qualitative behavior of the reducedorder<br />

model to develop some simplification of this control<br />

problem. As shown in Figure 3, the eigenvalue spectrum<br />

of the unactuated POD model indicates that most of the<br />

energy is contained in the first two POD modes. Moreover,<br />

the q 1 and q 2 , corresponding to these modes, have shedding<br />

frequency (f s ) while the second harmonic (2f s ) of vortex<br />

shedding frequency appears in q 3 and q 4 . It is also interesting<br />

to note that <strong>for</strong> the flow past a circular cylinder, the<br />

dominant frequency, i.e. vortex shedding frequency, and its<br />

odd harmonics (3f s , 5f s ) appear in the lift coefficient while<br />

even harmonics (2f s , 4f s ) appear in the drag coefficient.<br />

From the control perspective, if we are able to control the<br />

states q, we can reduce the fluctuating <strong>for</strong>ces, i.e. C L , on<br />

the cylinder. Thus, in the present setting, the optimization<br />

problem can be redefined as<br />

min<br />

∫ ∞<br />

0<br />

q T (t)Qq(t) + Rγc 2 (t) dt.<br />

In general, we define the state weighting matrix Q M×M and<br />

the control ef<strong>for</strong>t weighting matrix R Mc×M c<br />

. In the current<br />

study, we choose M = 12, M c = 1, Q = I, and R = 10.<br />

We compute the appropriate gains <strong>using</strong> LQR function in<br />

MATLAB and observe the system response. Feeding back<br />

the optimal gains, all the states decay to zero. In Figure 7(a),<br />

we plot only the first state (q 1 ) to show its response. We also<br />

observe that as the states approach zero, the control input<br />

γ c also approaches zero as shown in Figure 7(b). However,<br />

this procedure requires in<strong>for</strong>mation of all the states. From<br />

application point of view, we need observers to estimate<br />

all the states to achieve the objective, thus requiring greater<br />

ef<strong>for</strong>t and more approximation. We have also noted that states<br />

corresponding to higher modes have higher harmonics of<br />

vortex shedding which requires sensing the flow at higher<br />

frequency. In other words, <strong>using</strong> such an approach requires<br />

computation of all the states which may not be possible in<br />

a realistic flow field.<br />

B. Feeding back the Dominant State<br />

As discussed be<strong>for</strong>e, the first pair of states (q 1 and q 2 )<br />

corresponds to the dominant frequency, i.e. f s , while subsequent<br />

pairs have harmonics of the fundamental frequency.<br />

The objective here is to feedback only one state and observe<br />

its effect on the control input and the response of the system.<br />

We feedback only q 1 with its gain and observe the response<br />

of the system with different values of R. We study three<br />

4<br />

3<br />

2<br />

1<br />

q 1 0<br />

γ c<br />

−1<br />

−2<br />

−3<br />

−4<br />

0 50 100 150 200<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

−0.4<br />

−0.6<br />

Time<br />

(a) q 1<br />

−0.8<br />

0 50 100 150 200<br />

Fig. 7.<br />

Time<br />

(b) γ c<br />

Time histories of q 1 and γ c with full-state feedback.<br />

cases involving q 1 and with varying |Q|<br />

R<br />

as shown in Table<br />

I. We observe that feeding back only the dominant state (q 1 )<br />

in the system can stabilize it and the response the states<br />

decay to zero <strong>for</strong> all the three cases as shown in Figures<br />

8 and 9. Based on the magnitude of |Q|<br />

R<br />

, the response is<br />

different <strong>for</strong> each case, however, qualitatively the behavior<br />

shows a similar trend. We also compute the settling time and<br />

maximum control input required in each case. From Figures<br />

TABLE I<br />

SYSTEM RESPONSE.<br />

|Q|<br />

Case R Gain Settling time max γ c<br />

A 1 0.9348 50 2.25<br />

B 0.1 0.3253 120 1.1<br />

C 0.01 0.1105 250 1.05<br />

8 and 9, we observe that <strong>for</strong> Case A, the settling time is 50<br />

time units. As we decrease the Q11<br />

R<br />

, the settling time increase<br />

<strong>for</strong> Cases B and C to 120 and 250, respectively. Likewise,<br />

in Figure 10, we observe that the gain associated with the<br />

first mode is high <strong>for</strong> Case A as compared to Cases B and C<br />

which is obvious due to increase in the control ef<strong>for</strong>t. Note<br />

that γ c → 0 in each case. Thus, given the control ef<strong>for</strong>t we<br />

can suppress the response of the system which eventually<br />

leads to the suppression of fluctuating <strong>for</strong>ces on the cylinder.<br />

Proceedings of International Bhurban Conference on Applied Sciences & Technology,<br />

Islamabad, Pakistan, 10 - 13 January, 2011<br />

83


q1(t)<br />

3<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

−3<br />

0 50 100 150 200<br />

Time<br />

(a) Case A<br />

q2(t)<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

−2<br />

−2.5<br />

−3<br />

0 50 100 150 200<br />

Time<br />

(a) Case A<br />

3<br />

3<br />

2<br />

2<br />

1<br />

1<br />

q1(t)<br />

0<br />

q2(t)<br />

0<br />

−1<br />

−1<br />

−2<br />

−2<br />

−3<br />

0 50 100 150 200<br />

Time<br />

(b) Case B<br />

−3<br />

0 50 100 150 200<br />

Time<br />

(b) Case B<br />

q1(t)<br />

3<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

−3<br />

0 50 100 150 200<br />

Time<br />

(c) Case C<br />

Fig. 8. Time histories of q 1 with feeding back dominant state (q 1 ).<br />

q2(t)<br />

4<br />

3<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

−3<br />

0 50 100 150 200<br />

Time<br />

(c) Case C<br />

Fig. 9. Time histories of q 2 with feeding back dominant state (q 1 ).<br />

C. Feeding back the Harmonics<br />

In an attempt to analyze the effect of states other than<br />

q 1 , we feedback higher harmonics such as q 3 to the system<br />

along with its corresponding gain. It is important to note<br />

that the dominant frequency in q 3 is 2f s . Unlike q 1 , the<br />

system diverges as we integrate <strong>for</strong> a long time, as shown in<br />

Figure 11. We suggest that divergence in the system is due<br />

to absence of fundamental frequency in the feedback signal.<br />

A similar behavior is observed while feeding back higher<br />

harmonics. It may be related to the fact that these states are<br />

less dominant and does not correspond to the fundamental<br />

frequency of the system which appears in the lift coefficient.<br />

Thus, feedback of the dominant state may be sufficient to<br />

control the system dynamics and we may not require to<br />

estimate all the states in the system.<br />

CONCLUSIONS<br />

Using the snapshots of the flow field past a circular<br />

cylinder, we developed a low-dimensional model <strong>using</strong> POD-<br />

Galerkin expansion approach. Suction is used as the mechanism<br />

to control the flow field around a cylinder. Using the<br />

DNS of the controlled flow field, we developed a suitable<br />

control function that satisfies the inhomogeneous boundary<br />

condition on the cylinder surface. We substituted the modified<br />

velocity expansion into the Navier-Stokes equations and<br />

projected it onto the velocity POD modes. We linearized the<br />

system about the mean flow and reduced the model to an LTI<br />

Proceedings of International Bhurban Conference on Applied Sciences & Technology,<br />

Islamabad, Pakistan, 10 - 13 January, 2011<br />

84


γc<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

−2<br />

−2.5<br />

0 50 100 150 200<br />

Time<br />

1.2<br />

(a) Case A<br />

4<br />

3<br />

2<br />

1<br />

q 1 0<br />

−1<br />

−2<br />

−3<br />

−4<br />

0 50 100 150 200<br />

1<br />

Time<br />

(a) q 1<br />

1<br />

γc<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

γ c<br />

0.5<br />

0<br />

0<br />

−0.2<br />

−0.5<br />

−0.4<br />

−0.6<br />

0 50 100 150 200<br />

Time<br />

1.2<br />

1<br />

(b) Case B<br />

−1<br />

0 50 100 150 200<br />

Time<br />

(b) γ c<br />

Fig. 11. Time histories of q 1 and γ c with feeding back higher harmonic<br />

(q 3 ) only.<br />

0.8<br />

γc<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

0 50 100 150 200<br />

Time<br />

(c) Case C<br />

Fig. 10. Time histories of γ c with feeding back dominant state (q 1 ).<br />

system. We applied optimal control on the LTI system with<br />

the objective to reduce the fluctuating <strong>for</strong>ces on the cylinder.<br />

We studied different cases by varying Q R<br />

ratios and studied<br />

the system response.<br />

Note that our control is defined by feeding back the<br />

dominant state of the model. We observed that the dominant<br />

state is proportional to the lift coefficient. The latter can<br />

be estimated by placing sensors on the cylinder surface,<br />

thus with appropriate gains, the feedback control system is<br />

practically implementable. We showed that feedback higher<br />

harmonics in fact fails to stabilize the system.<br />

ACKNOWLEDGMENTS<br />

This research was supported in part by Air Force Office of<br />

Scientific Research grants FA9550-07-1-0273 and FA9550-<br />

10-1-0201 and by the Environmental Security Technology<br />

Certification Program (ESTCP) under a subcontract from the<br />

United Technologies.<br />

REFERENCES<br />

[1] K. Ito and S. Ravindran, “Reduced basis method <strong>for</strong> flow control,”<br />

1996, technical Report CRSC-TR96-25, Center <strong>for</strong> Research in<br />

Scientific Computation, North Carolina State University, Raleigh,<br />

NC. [<strong>On</strong>line]. Available: citeseer.ist.psu.edu/article/ito96reduced.html<br />

[2] R. Bishop and A. Hassan, “The lift and drag <strong>for</strong>ces on a circular<br />

cylinder in flowing fluid,” Proceedings of the Royal Society Series A,<br />

vol. 277, pp. 32–50, 1963.<br />

[3] R. Hartlen and I. Currie, “Lift-oscillator model of vortex-induced<br />

vibration,” ASCE Journal of Engineering Mechanics, vol. 96, pp. 577–<br />

591, 1970.<br />

[4] W. Iwan and R. Blevins, “A model <strong>for</strong> vortex-induced oscillation of<br />

structures,” Journal of Applied Mechanics, vol. 41, no. 3, pp. 581–586,<br />

1974.<br />

[5] R. Landl, “A mathematical model <strong>for</strong> vortex-excited vibration of cable<br />

suspensions,” Journal of Sound and Vibration, vol. 42, no. 2, pp. 219–<br />

234, 1975.<br />

[6] R. Skop and O. Griffin, “<strong>On</strong> a theory <strong>for</strong> the vortex-excited oscillations<br />

of flexible cylindrical structures,” Journal of Sound and Vibration,<br />

vol. 41, no. 3, pp. 263–274, 1975.<br />

[7] S. Krenk and S. Nielsen, “Energy balanced double oscillator model <strong>for</strong><br />

vortex-induced vibrations,” ASCE Journal of Engineering Mechanics,<br />

vol. 125, no. 3, pp. 263–271, 1999.<br />

Proceedings of International Bhurban Conference on Applied Sciences & Technology,<br />

Islamabad, Pakistan, 10 - 13 January, 2011<br />

85


[8] O. A. Marzouk, A. H. Nayfeh, I. Akhtar, and H. N. Arafat, “<strong>Model</strong>ing<br />

steady-state and transient <strong>for</strong>ces on a cylinder,” Journal of Vibration<br />

and <strong>Control</strong>, vol. 13, no. 7, pp. 1065–1091, 2007.<br />

[9] I. Akhtar, O. A. Marzouk, and A. H. Nayfeh, “A van der Pol-Duffing<br />

oscillator model of hydrodynamic <strong>for</strong>ces on canonical structures,”<br />

Journal of Computational and Nonlinear Dynamics, vol. 4, no. 4, p.<br />

041006, 2009.<br />

[10] G. Berkooz, P. Holmes, and J. L. Lumley, “The proper orthogonal<br />

decomposition in the analysis of turbulent flows,” Annual Review of<br />

<strong>Fluid</strong> Mechanics, vol. 53, pp. 321–575, 1993.<br />

[11] P. Holmes, J. L. Lumley, and G. Berkooz, Turbulence, Coherent<br />

Structures, Dynamical Systems and Symmetry. Cambridge, UK:<br />

Cambridge University Press, 1996.<br />

[12] W. R. Graham, J. Peraire, and K. Y. Tang, “Optimal control of vortex<br />

shedding <strong>using</strong> low-order models. Part I - Open-loop model development,”<br />

International Journal <strong>for</strong> Numerical Methods in Engineering,<br />

vol. 44, pp. 945–972, 1999.<br />

[13] S. N. Singh, J. H. Myatt, G. A. Addington, S. Banda, and J. K. Hall,<br />

“Optimal feedback control of vortex shedding <strong>using</strong> proper orthogonal<br />

decomposition models,” Journal of <strong>Fluid</strong>s Engineering, vol. 123, pp.<br />

612–618, 2001.<br />

[14] M. Bergmann, L. Cordier, and J. P. Brancher, “Optimal rotary control<br />

of the cylinder wake <strong>using</strong> POD reduced-order model,” Physics of<br />

<strong>Fluid</strong>s, vol. 17, no. 9, pp. 097 101–20, 2005.<br />

[15] S. Siegel, K. Cohen, and T. McLaughlin, “Numerical simulations of<br />

a feedback-controlled circular cylinder wake,” AIAA Journal, vol. 44,<br />

no. 6, pp. 1266–1276, 2006.<br />

[16] M. Bergmann and L. Cordier, “Optimal rotary control of the cylinder<br />

wake in the laminar regime by trust-region methods and POD reducedorder<br />

model,” Journal of Computational Physics, vol. 227, no. 16, pp.<br />

7813–7840, 2008.<br />

[17] I. Akhtar, “Parallel simulations, reduced-order modeling, and feedback<br />

control of vortex shedding <strong>using</strong> fluidic actuators,” Ph.D. dissertation,<br />

Virginia Tech, Blacksburg, VA, 2008.<br />

[18] I. Akhtar, A. H. Nayfeh, and C. J. Ribbens, “<strong>On</strong> the stability and<br />

extension of reduced-order Galerkin models in incompressible flows:<br />

A numerical study of vortex shedding,” Theoretical and Computational<br />

<strong>Fluid</strong> Dynamics, vol. 23, no. 3, pp. 213–237, 2009.<br />

[19] I. Akhtar and A. H. Nayfeh, “<strong>Model</strong> based control of vortex shedding<br />

<strong>using</strong> fluidic actuators,” Journal of Computational and Nonlinear<br />

Dynamics, vol. 5, no. 4, p. 041015, 2010.<br />

[20] H. P. Bakewell and J. L. Lumley, “Viscous sublayer and adjacent wall<br />

region in turbulent pipe flow,” The Physics of <strong>Fluid</strong>s, vol. 10, no. 9,<br />

pp. 1880–1889, 1967.<br />

[21] L. Sirovich, “Turbulence and the dynamics of coherent structures,”<br />

Quarterly of Applied Mathematics, vol. 45, pp. 561–590, 1987.<br />

[22] C. Beattie, J. Borggaard, S. Guercin, and T. Iliescu, “A domain<br />

decomposition approach to POD,” in Processdings of the 45th IEEE<br />

Conference on Decision and <strong>Control</strong>. pp. 6750-6756, 2006.<br />

[23] A. E. Deane, I. G. Kevrekidis, G. E. Karniadakis, and S. A. Orsag,<br />

“Low-dimensional models <strong>for</strong> complex geometry flows: Application<br />

to grooved channels and circular cylinder,” Physics of <strong>Fluid</strong>s A, vol. 3,<br />

no. 10, pp. 2337–2354, 1991.<br />

[24] B. R. Noack, P. Papas, and P. A. Monkewitz, “The need <strong>for</strong> a pressureterm<br />

representation in empirical Galerkin models of incompressible<br />

shear flows,” Journal of <strong>Fluid</strong> Mechanics, vol. 523, pp. 339–365, 2005.<br />

[25] X. Ma and G. Karniadakis, “A low-dimensional model <strong>for</strong> simulating<br />

three-dimensional cylinder flow,” Journal of <strong>Fluid</strong> Mechanics, vol.<br />

458, pp. 181–190, 2002.<br />

[26] I. Akhtar, A. H. Nayfeh, and C. J. Ribbens, “<strong>On</strong> controlling the bluff<br />

body wake <strong>using</strong> a reduced-order model,” in Proceedings of the 4th<br />

Flow <strong>Control</strong> Conference, Seattle, WA. AIAA Paper No. 2008-4189,<br />

2008.<br />

[27] V. I. Arnold, Geometric Methods in the Theory of Ordinary Differential<br />

Equations. New York: Springer-Verlag, 1988, ch. 3.<br />

Proceedings of International Bhurban Conference on Applied Sciences & Technology,<br />

Islamabad, Pakistan, 10 - 13 January, 2011 86

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!