02.12.2012 Views

Numerical Simulation of a Cylinder in Uniform Flow: Application of a ...

Numerical Simulation of a Cylinder in Uniform Flow: Application of a ...

Numerical Simulation of a Cylinder in Uniform Flow: Application of a ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

JOURNAL OF COMPUTATIONAL PHYSICS 123, 450–465 (1996)<br />

ARTICLE NO. 0036<br />

<strong>Numerical</strong> <strong>Simulation</strong> <strong>of</strong> a <strong>Cyl<strong>in</strong>der</strong> <strong>in</strong> <strong>Uniform</strong> <strong>Flow</strong>: <strong>Application</strong> <strong>of</strong><br />

a Virtual Boundary Method<br />

E. M. SAIKI AND S. BIRINGEN<br />

Department <strong>of</strong> Aerospace Eng<strong>in</strong>eer<strong>in</strong>g Sciences, University <strong>of</strong> Colorado, Boulder, Colorado<br />

Received July 12, 1994; revised July 10, 1995<br />

In this study, a virtual boundary technique is applied to the numer-<br />

<strong>in</strong> its ability to model the material properties <strong>of</strong> the body<br />

and movement <strong>of</strong> the boundaries. However, numerical<br />

ical simulation <strong>of</strong> stationary and mov<strong>in</strong>g cyl<strong>in</strong>ders <strong>in</strong> uniform flow. stiffness <strong>of</strong> most mov<strong>in</strong>g boundary problems restricts the<br />

This approach readily allows the imposition <strong>of</strong> a no-slip boundary<br />

with<strong>in</strong> the flow field by a feedback forc<strong>in</strong>g term added to the momen-<br />

tum equations. In the present work, this technique is used with a<br />

high-order f<strong>in</strong>ite difference method, effectively elim<strong>in</strong>at<strong>in</strong>g spurious<br />

oscillations caused by the feedback forc<strong>in</strong>g when used with spec-<br />

explicit def<strong>in</strong>ition <strong>of</strong> the forc<strong>in</strong>g term <strong>in</strong> Pesk<strong>in</strong>’s method<br />

to small time steps (Tu and Pesk<strong>in</strong> [32]). This method has<br />

been expanded and implemented <strong>in</strong> a number <strong>of</strong> other<br />

problems model<strong>in</strong>g suspended particulates (Fogelson and<br />

trally discretized flow solvers. Very good agreement is found be- Pesk<strong>in</strong> [12]) and the <strong>in</strong>ner ear (Beyer [3]).<br />

tween the present calculations and previous computational and<br />

experimental results for steady and time-dependent flow at low<br />

Reynolds numbers. � 1996 Academic Press, Inc.<br />

In a related, yet <strong>in</strong>dependent, study, Goldste<strong>in</strong> et al. [15,<br />

16] developed a virtual boundary method based on the<br />

<strong>in</strong>itial work <strong>of</strong> Sirovich [28, 29] which employs a forc<strong>in</strong>g<br />

term governed by a feedback loop. Us<strong>in</strong>g a spectral<br />

INTRODUCTION<br />

method, they applied this procedure to <strong>in</strong>vestigate the effects<br />

<strong>of</strong> riblets on turbulent channel flow and to flow be-<br />

The fundamental fluid dynamics problem <strong>of</strong> a circular tween concentric cyl<strong>in</strong>ders. They noted that the forc<strong>in</strong>g<br />

cyl<strong>in</strong>der <strong>in</strong> uniform flow has been exam<strong>in</strong>ed extensively function generated constant low amplitude, high frequency<br />

<strong>in</strong> both computational and experimental studies and is oscillations which they were able to control by numerical<br />

considered a str<strong>in</strong>gent test for flow solvers. The difficulty low-pass filters and/or the <strong>in</strong>troduction <strong>of</strong> particular flow<br />

which accompanies the computational approach to this fields <strong>in</strong>side <strong>of</strong> the body. Their simulations were not notice-<br />

problem by f<strong>in</strong>ite differences or spectral methods lies <strong>in</strong> ably affected by these spurious signals, but such numerical<br />

the representation <strong>of</strong> the cyl<strong>in</strong>der geometry to allow for oscillations may become <strong>of</strong> concern when one calculates<br />

an accurate application <strong>of</strong> these numerical <strong>in</strong>tegration the evolution <strong>of</strong> a forced disturbance wave as <strong>in</strong> the simula-<br />

methods. The use <strong>of</strong> coord<strong>in</strong>ate transformations and map- tion <strong>of</strong> flow <strong>in</strong>stability and transition.<br />

p<strong>in</strong>g techniques is possible but requires a highly accurate In the present study, we use the method developed by<br />

way <strong>of</strong> calculat<strong>in</strong>g the transformation Jacobians. F<strong>in</strong>ite ele- Goldste<strong>in</strong> et al. [16] to simulate stationary, rotat<strong>in</strong>g, and<br />

ment methods (Gresho et al.[17]; Engelman and Jam<strong>in</strong>ia oscillat<strong>in</strong>g cyl<strong>in</strong>ders <strong>in</strong> uniform flow at low Reynolds num-<br />

[11]; Karniadakis and Triantafyllou [21]) and conformal bers (Re � 400) allow<strong>in</strong>g the assessment <strong>of</strong> the virtual<br />

transformations (Jordan and Fromm [20]; Braza et al. [4]; boundary technique to model a body <strong>in</strong> an unsteady flow<br />

Badr and Dennis [1]) have been successfully used for this field. In the present solution procedure, high-order f<strong>in</strong>ite<br />

problem. As an alternative to the use <strong>of</strong> generalized coordi- differences are implemented <strong>in</strong> order to suppress the nunates<br />

and coord<strong>in</strong>ate transformations for f<strong>in</strong>ite difference merical oscillations caused by the forc<strong>in</strong>g function ob-<br />

and spectral methods, Pesk<strong>in</strong> [23] developed a method served <strong>in</strong> the Chebyshev spectral method <strong>of</strong> Goldste<strong>in</strong> et<br />

which represents a body with<strong>in</strong> a flow field via a forc<strong>in</strong>g<br />

term added to the govern<strong>in</strong>g equations. When applied at<br />

al. [16].<br />

certa<strong>in</strong> po<strong>in</strong>ts <strong>in</strong> the flow, this forc<strong>in</strong>g term simulates the<br />

effect <strong>of</strong> the body on the flow, allow<strong>in</strong>g for the model<strong>in</strong>g<br />

COMPUTATIONAL METHOD<br />

<strong>of</strong> a boundary <strong>of</strong> any shape with<strong>in</strong> a Cartesian computa- The numerical model <strong>in</strong>tegrates the two-dimensional,<br />

tional box without the necessity <strong>of</strong> mapp<strong>in</strong>g. Pesk<strong>in</strong> [23, time-dependent, <strong>in</strong>compressible, Navier-Stokes, and con-<br />

24] successfully implemented this method (immersed t<strong>in</strong>uity equations nondimensionalized by the diameter <strong>of</strong><br />

boundary technique) to model mov<strong>in</strong>g boundaries <strong>in</strong> heart the cyl<strong>in</strong>der, D, and the free-stream velocity, U�,ona valve simulations. The ma<strong>in</strong> advantage <strong>of</strong> the scheme lies staggered mesh by a time-splitt<strong>in</strong>g method. The normal<br />

0021-9991/96 $18.00<br />

Copyright © 1996 by Academic Press, Inc.<br />

All rights <strong>of</strong> reproduction <strong>in</strong> any form reserved.<br />

450


VIRTUAL BOUNDARY METHOD 451<br />

FIG. 1. Time distribution <strong>of</strong> the l 2-norm error <strong>in</strong> the no-slip condition imposed by the virtual boundary method def<strong>in</strong><strong>in</strong>g a cyl<strong>in</strong>der <strong>in</strong> uniform<br />

flow at Re � 25. Comparison between solutions with vary<strong>in</strong>g �/�: (a) over total run time; (b) closeup <strong>of</strong> first 1000 time steps; and (c) closeup <strong>of</strong><br />

last 1000 time steps. � ��400000, � ��600, ———; � ��4000, � ��60, –––; � ��400, � ��6, –---–---.<br />

diffusion terms are advanced implicitly by the Crank– boundary, uniform flow conditions are assumed, i.e., u �<br />

Nicolson scheme and either the explicit third-order com- 1 and v � 0. At the outflow, boundary conditions are<br />

pact Runge–Kutta or Adams–Bashforth methods are prescribed to ensure that wave-like disturbances (gener-<br />

applied to the rema<strong>in</strong><strong>in</strong>g terms (Streett and Hussa<strong>in</strong>i ated by vortex shedd<strong>in</strong>g) <strong>in</strong> the high Reynolds number<br />

[30]). The equations are discretized spatially <strong>in</strong> the normal cases leave the computational doma<strong>in</strong> without reflection.<br />

( y) and streamwise (x) directions by fourth-order central This was accomplished by append<strong>in</strong>g a ‘‘buffer doma<strong>in</strong>’’<br />

f<strong>in</strong>ite differences. The pressure Poisson equation is evalu- to the physical doma<strong>in</strong> (the length <strong>of</strong> the buffer doma<strong>in</strong><br />

ated by the tensor product method (Huser and Bir- was about 20–30% <strong>of</strong> the physical doma<strong>in</strong>) <strong>in</strong> which the<br />

<strong>in</strong>gen [19]). govern<strong>in</strong>g equations were modified by reduc<strong>in</strong>g the stream-<br />

At the upper and lower boundaries, we impose shear wise viscous terms and the right-hand side <strong>of</strong> the pressure<br />

free conditions, i.e., �u/�y � 0 and v � 0 and at the <strong>in</strong>flow Poisson equation to zero at the outflow boundary us<strong>in</strong>g a


452 SAIKI AND BIRINGEN<br />

FIG. 2. Re � 550, t � 3: streamwise velocity. (a) Chebyshev; (b) f<strong>in</strong>ite differences.<br />

smooth coefficient function. Previous numerical experi- body moves, i.e., v � 0, then the position <strong>of</strong> the boundary<br />

ments have <strong>in</strong>cluded rigorous test<strong>in</strong>g <strong>of</strong> this technique, po<strong>in</strong>ts at each time step is computed by <strong>in</strong>tegration <strong>of</strong><br />

verify<strong>in</strong>g its suitability for use <strong>in</strong> both high and low ampli- v � dxs/dt. The negative constants � and � are determ<strong>in</strong>ed<br />

tude wave propagation problems (Streett and Macaraeg by observ<strong>in</strong>g the response <strong>of</strong> U once F is applied; � pro-<br />

[31]; Danabasoglu [9]; Danabasoglu et al. [10]; Saiki et duces the natural oscillation frequency <strong>of</strong> the response,<br />

al. [26]). while � dampens the oscillation <strong>of</strong> the response. For un-<br />

In implement<strong>in</strong>g the method <strong>of</strong> Goldste<strong>in</strong> et al. [16] to steady flows, � must produce a response with a natural<br />

the present calculations <strong>of</strong> flow over a cyl<strong>in</strong>der, the no- frequency greater than the highest frequencies present <strong>in</strong><br />

slip boundary <strong>of</strong> the cyl<strong>in</strong>der surface was represented by the flow so that F can respond correctly to the chang<strong>in</strong>g<br />

a feedback forc<strong>in</strong>g function added to the momentum equa- flow field. The oscillatory nature <strong>of</strong> the boundary does not<br />

tions. This feedback function effectively br<strong>in</strong>gs the fluid affect the overall steady state <strong>of</strong> the flow field. However,<br />

velocity to zero at the desired po<strong>in</strong>ts <strong>in</strong> the flow which for each comb<strong>in</strong>ation <strong>of</strong> � and � a different forc<strong>in</strong>g function<br />

def<strong>in</strong>e the no-slip boundary and can be expressed as is added to the right-hand side <strong>of</strong> the momentum equations,<br />

F(xs, t) � ��<br />

therefore a set <strong>of</strong> similar but slightly different flow fields<br />

are obta<strong>in</strong>ed for each <strong>of</strong> the solutions. For example, com-<br />

t<br />

0 (U(xs, t) � v(xs, t)) dt<br />

� �(U(xs, t) � v(xs, t)).<br />

(1) parison <strong>of</strong> the cyl<strong>in</strong>der cases presented below revealed<br />

variations <strong>of</strong> 1–2% <strong>in</strong> the geometrical parameters <strong>of</strong> the<br />

wake.<br />

Here, F is the external force imposed at the discrete surface To exam<strong>in</strong>e the response <strong>of</strong> the virtual boundary to<br />

po<strong>in</strong>ts def<strong>in</strong>ed by xs, and U is the fluid velocity at these different �/� comb<strong>in</strong>ations, a stationary cyl<strong>in</strong>der at Re �<br />

surface po<strong>in</strong>ts. The velocity <strong>of</strong> the body itself is controlled 25 was modeled and the l2-norm error <strong>of</strong> the velocities at<br />

by specify<strong>in</strong>g v � (ub, vb) at the boundary po<strong>in</strong>ts. If the the boundary po<strong>in</strong>ts with respect to the no-slip boundary


VIRTUAL BOUNDARY METHOD 453<br />

FIG. 3. Re � 550: normal pr<strong>of</strong>iles <strong>of</strong> streamwise velocity <strong>in</strong> the vic<strong>in</strong>ity <strong>of</strong> the cyl<strong>in</strong>der. (a) x � 2.23750; (b) x � 2.2875; (c) x � 2.3125;<br />

(d) x � 2.3650. Chebyshev, ———; f<strong>in</strong>ite differences –––.<br />

condition was tracked <strong>in</strong> time. The l2-norm <strong>in</strong> the stream- The time step restrict<strong>in</strong>g this method is based upon the<br />

wise velocity is def<strong>in</strong>ed as<br />

values <strong>of</strong> � and � and the explicit time <strong>in</strong>tegration imple-<br />

mented <strong>in</strong> the flow solver. For the Adams–Bashforth<br />

method, Goldste<strong>in</strong> et al. [16] determ<strong>in</strong>ed the follow<strong>in</strong>g<br />

l 2-norm �� 1<br />

n b � n b<br />

i�1<br />

(u b) 2 i , (2) expression for the time step<br />

where nb is the number <strong>of</strong> virtual boundary po<strong>in</strong>ts. The �t � �� � �(�2 results <strong>of</strong> this analysis are presented <strong>in</strong> Fig. 1. With the<br />

application <strong>of</strong> the higher values <strong>of</strong> these parameters, the<br />

� 2�k)<br />

,<br />

�<br />

(3)<br />

no-slip boundary condition is quickly atta<strong>in</strong>ed. At the end where k is a problem dependent constant <strong>of</strong> order one. A<br />

<strong>of</strong> the time period considered, the l2-norm correspond<strong>in</strong>g similar restriction was observed <strong>in</strong> the current study when<br />

to the lower �/� values is two orders <strong>of</strong> magnitude higher the Runge–Kutta method was used.<br />

than the other two cases; however, it cont<strong>in</strong>ues to decay. In order to represent the body boundary <strong>in</strong> the flow<br />

These test cases suggest that higher values <strong>of</strong> the coeffi- field, Goldste<strong>in</strong> et al. [16] def<strong>in</strong>ed the boundary on<br />

cients allow the method to respond faster to any unsteadi- po<strong>in</strong>ts which co<strong>in</strong>cided with the computational grid nodes,<br />

ness <strong>in</strong> the flow field and act more efficiently <strong>in</strong> re<strong>in</strong>forc<strong>in</strong>g whereas Pesk<strong>in</strong> [23, 24] and Goldste<strong>in</strong> et al. [15] def<strong>in</strong>ed<br />

the no-slip conditions. These computations were con- the body boundary <strong>in</strong> a manner <strong>in</strong>dependent <strong>of</strong> the<br />

ducted us<strong>in</strong>g nb � 1441 boundary po<strong>in</strong>ts; further <strong>in</strong>creases computational grid. Pesk<strong>in</strong> [23] utilized a first-order co-<br />

or decreases <strong>in</strong> nb with the same values <strong>of</strong> � and � yielded s<strong>in</strong>e function to <strong>in</strong>terpolate and extrapolate <strong>in</strong>formation<br />

similar results. between the immersed boundary and the grid. More


454 SAIKI AND BIRINGEN<br />

FIG. 4. Superposition <strong>of</strong> the virtual boundary po<strong>in</strong>ts on the computa-<br />

tional grid: (a) full cyl<strong>in</strong>der body; (b) upper left portion <strong>of</strong> the cyl<strong>in</strong>der.<br />

recently, Beyer [3] developed a second-order accurate<br />

representation <strong>of</strong> the immersed boundary <strong>in</strong> applications<br />

<strong>of</strong> Pesk<strong>in</strong>’s method. Goldste<strong>in</strong> et al. [15] implemented<br />

highly accurate spectral <strong>in</strong>terpolation <strong>of</strong> the velocities<br />

from the grid po<strong>in</strong>ts to the virtual boundary po<strong>in</strong>ts and<br />

applied l<strong>in</strong>ear <strong>in</strong>terpolation to distribute the effect <strong>of</strong><br />

the forc<strong>in</strong>g term to the grid nodes.<br />

In the current study, the fluid velocities are <strong>in</strong>terpolated<br />

to a virtual boundary po<strong>in</strong>t, (xs, ys), from the four sur-<br />

round<strong>in</strong>g grid po<strong>in</strong>ts denoted by the <strong>in</strong>dices (i, j),<br />

(i � 1, j), (i, j � 1), and (i � 1, j � 1), us<strong>in</strong>g bil<strong>in</strong>ear <strong>in</strong>ter-<br />

polation,<br />

i�1, j�1<br />

U(xs) � � Di, j(xs)Ui, j, (4)<br />

i, j<br />

where<br />

In Eq. (5),<br />

and<br />

D i, j(x s) � d(x s � x i)d( y s � y j). (5)<br />

d(x s � x i) � (x s � x i�1)<br />

(x i � x i�1)<br />

d(x s � x i) � (x s � x i�1)<br />

(x i � x i�1)<br />

if x i � x s<br />

if x i � x s<br />

(6a)<br />

(6b)<br />

d(x s � x i) � 1 ifx i�x s. (6c)<br />

The effect <strong>of</strong> the virtual boundary force is extrapolated<br />

back to the grid po<strong>in</strong>ts by area-weighted averages,<br />

F i, j � 1<br />

N b � N b<br />

n�1<br />

D i, j(x s)F n(x s), (7)<br />

where Nb is the number <strong>of</strong> virtual boundary po<strong>in</strong>ts which<br />

affect the (i, j)th grid po<strong>in</strong>t. This method <strong>of</strong> spread<strong>in</strong>g the<br />

boundary forces results <strong>in</strong> an effective boundary thickness<br />

on the order <strong>of</strong> one grid cell, i.e., O(�x, �y). The above<br />

<strong>in</strong>terpolation/extrapolation scheme is first-order accurate,<br />

similar to the delta function representation <strong>of</strong> Pesk<strong>in</strong> [23].<br />

The low order accuracy <strong>of</strong> this operation <strong>in</strong>fluences ma<strong>in</strong>ly<br />

the flow field <strong>in</strong> the immediate vic<strong>in</strong>ity <strong>of</strong> the cyl<strong>in</strong>der;<br />

however, large scale features are successfully captured by<br />

this method.<br />

Both Pesk<strong>in</strong> [23, 24] and Goldste<strong>in</strong> et al. [15, 16] imposed<br />

the forc<strong>in</strong>g term only at po<strong>in</strong>ts which def<strong>in</strong>ed the boundary,<br />

thus allow<strong>in</strong>g fluid motion <strong>in</strong>side the body. For Pesk<strong>in</strong>’s<br />

[23, 24] work this behavior is desirable s<strong>in</strong>ce his calculations<br />

concern blood flow <strong>in</strong>side <strong>of</strong> the heart and the external<br />

flow field is ignored. Goldste<strong>in</strong> et al. [15, 16] <strong>in</strong>vestigated<br />

the effect <strong>of</strong> solid bodies placed with<strong>in</strong> a flow field which<br />

physically do not permit flow <strong>in</strong>side the boundary. Conse-<br />

quently, the flow fields which were numerically allowed to<br />

develop <strong>in</strong> such boundaries were unphysical; however, the<br />

<strong>in</strong>ternal flow field was used as a smooth<strong>in</strong>g device to attenu-<br />

ate spatial oscillations generated by the method (Goldste<strong>in</strong><br />

et al. [15, 16]). In the present computations, some <strong>of</strong> the test<br />

cases (<strong>in</strong> particular, the Re � 550 case presented below)<br />

converged to an <strong>in</strong>correct solution with the forc<strong>in</strong>g term<br />

imposed only on the boundary. This behavior was remedied<br />

by impos<strong>in</strong>g the forc<strong>in</strong>g function <strong>in</strong>side the boundary<br />

<strong>of</strong> the body, as well as on the boundary. This suggests that<br />

<strong>in</strong> implement<strong>in</strong>g the virtual boundary technique <strong>in</strong> solid<br />

body problems, where the solution is unknown, the forc<strong>in</strong>g


VIRTUAL BOUNDARY METHOD 455<br />

FIG. 5. Re � 25: (a) spanwise vorticity, dotted and solid l<strong>in</strong>es denote negative and positive levels, respectively; (b) streamfunction.<br />

term should be applied at the boundary and <strong>in</strong>terior po<strong>in</strong>ts the attenuation <strong>of</strong> the oscillations with the application <strong>of</strong><br />

<strong>of</strong> the body <strong>in</strong> order to converge to a correct solution. f<strong>in</strong>ite differences. Comparison <strong>of</strong> the streamwise pr<strong>of</strong>iles<br />

In an earlier work by the authors [27], the virtual bound- <strong>in</strong> the normal direction (Fig. 3) provide clear evidence that<br />

ary technique was used with a different flow solver imple- the oscillations are strongly damped.<br />

ment<strong>in</strong>g Chebyshev polynomials <strong>in</strong> the wall-normal direction.<br />

Due to the global nature <strong>of</strong> the Chebyshev<br />

RESULTS<br />

polynomials, nongrow<strong>in</strong>g, spatial oscillations <strong>in</strong> the normal<br />

and streamwise directions developed <strong>in</strong> the flow field when<br />

Stationary <strong>Cyl<strong>in</strong>der</strong> <strong>in</strong> <strong>Uniform</strong> <strong>Flow</strong><br />

the feedback forc<strong>in</strong>g function was applied. The virtual The cases <strong>in</strong>vestigat<strong>in</strong>g uniform flow over a stationary<br />

boundary acts as a discont<strong>in</strong>uity <strong>in</strong> the spectral representa- cyl<strong>in</strong>der exam<strong>in</strong>ed Reynolds numbers (Re � U�D/�) rangtion<br />

<strong>of</strong> the flow field, thus caus<strong>in</strong>g the oscillations to arise <strong>in</strong>g from Re � 25 to Re � 400. The mesh resolution varied<br />

(Gibb’s phenomenon). These oscillations were similar to from 267 � 147 to 436 � 147 for computational doma<strong>in</strong>s<br />

those also observed by Goldste<strong>in</strong> et al. [16] and did not rang<strong>in</strong>g from 20 � 10 to 34 � 10. As the Reynolds number<br />

appear to affect the flow field downstream <strong>of</strong> the body. In <strong>in</strong>creased, the length <strong>of</strong> the computational doma<strong>in</strong> was<br />

the present work, the application <strong>of</strong> a local discretization <strong>in</strong>creased <strong>in</strong> order to accommodate the stronger vortices<br />

scheme, i.e., f<strong>in</strong>ite differences, <strong>in</strong> the normal direction dras- which were shed from the cyl<strong>in</strong>der. These adjustments are<br />

tically reduced the amplitude <strong>of</strong> these spatial oscillations. reflected <strong>in</strong> the vary<strong>in</strong>g grid and doma<strong>in</strong> sizes cited above.<br />

The effect <strong>of</strong> these different discretization methods (i.e., Mesh stretch<strong>in</strong>g was employed <strong>in</strong> both directions with grid<br />

the Chebyshev and f<strong>in</strong>ite difference methods) is illustrated cluster<strong>in</strong>g near the body: the m<strong>in</strong>imum grid spac<strong>in</strong>g <strong>in</strong> the<br />

<strong>in</strong> results obta<strong>in</strong>ed from computations <strong>of</strong> startup flow over vic<strong>in</strong>ity <strong>of</strong> the cyl<strong>in</strong>der for all cases was �xm<strong>in</strong> ��ym<strong>in</strong> �<br />

a cyl<strong>in</strong>der at Re � 550 (Figs. 2 and 3). Contours <strong>of</strong> stream- 0.0375. The feedback forc<strong>in</strong>g coefficients were set to � �<br />

wise velocity reveal the spatial oscillations which arise due �400000 and � ��600, and the number <strong>of</strong> po<strong>in</strong>ts def<strong>in</strong><strong>in</strong>g<br />

to the Chebyshev discretization (Fig. 2a), and Fig. 2b shows the cyl<strong>in</strong>der was 1441. The distribution <strong>of</strong> po<strong>in</strong>ts def<strong>in</strong><strong>in</strong>g


456 SAIKI AND BIRINGEN<br />

FIG. 6. Re � 30 : (a) spanwise vorticity, dotted and solid l<strong>in</strong>es denote negative and positive levels, respectively; (b) streamfunction.<br />

TABLE I<br />

Comparison <strong>of</strong> the Wake Properties beh<strong>in</strong>d a Stationary <strong>Cyl<strong>in</strong>der</strong> with Experiments and Previous Computational Results<br />

for Re � 25 and Re � 30; Steady <strong>Flow</strong> Solutions<br />

Re � 25 Re � 30<br />

Coutanceau Coutanceau<br />

Properties <strong>of</strong> the wake beh<strong>in</strong>d a Present Gresho et al. & Bouard Clift et al. Present & Bouard Clift et al.<br />

stationary cyl<strong>in</strong>der results [17] a [6] b [5] b results [6] b [5] b<br />

Length <strong>of</strong> the separation bubble (L) 1.41 1.15 1.22 — 1.7 1.53 —<br />

x-coord<strong>in</strong>ate <strong>of</strong> the center <strong>of</strong> the<br />

vortex cores (a) 0.53 0.38 0.44 — 0.62 0.55 —<br />

y-distance between the vortex cores<br />

(b) 0.5 0.47 0.51 — 0.5625 0.54 —<br />

M<strong>in</strong>imum streamwise velocity on<br />

the axis <strong>of</strong> symmetry (u max) �0.064 �0.057 �0.057 — �0.08 �0.0743 —<br />

x-coord<strong>in</strong>ate <strong>of</strong> m<strong>in</strong>imum streamwise<br />

velocity on the axis <strong>of</strong> symmetry<br />

(d) 0.59 0.49 0.5 — 0.72 0.64 —<br />

Separation angle (�) 45� 45� 48� — 48� 50.1� —<br />

Drag coefficient (C d) 1.54 2.26 — 1.84 1.38 — 1.69<br />

a Computational study.<br />

b Experiment.


VIRTUAL BOUNDARY METHOD 457<br />

FIG. 7. Re � 50 : steady state solution. (a) spanwise vorticity, dotted and solid l<strong>in</strong>es denote negative and positive levels, respectively; (b) streamfunction.<br />

the virtual boundary superimposed onto the computational At low Reynolds numbers, (4.5 � Re � 35), experiments<br />

grid is presented <strong>in</strong> Fig. 4. In this particular case, we chose reveal an attached, steady, symmetric, recirculat<strong>in</strong>g bubble<br />

to impose the virtual boundary only on po<strong>in</strong>ts def<strong>in</strong><strong>in</strong>g the which develops downstream <strong>of</strong> the cyl<strong>in</strong>der (Coutanceau<br />

boundary <strong>of</strong> the cyl<strong>in</strong>der; similar results were obta<strong>in</strong>ed and Defaye [7]). In the present simulations this behavior<br />

when po<strong>in</strong>ts were imposed <strong>in</strong> the <strong>in</strong>terior. The velocity is clearly observed <strong>in</strong> the stream-function and spanwise<br />

components <strong>of</strong> the boundary po<strong>in</strong>ts, v, were set to zero. vorticity contours for Re � 25 and Re � 30, respectively<br />

FIG. 8. Re � 50 : unsteady solution. (a) spanwise voriticity, dotted and solid l<strong>in</strong>es denote negative and positive levels, respectively; (b) streamfunction.


458 SAIKI AND BIRINGEN<br />

TABLE II<br />

Comparison <strong>of</strong> Quantitative Data with Results from Experiments and Previous Computational Studies; Unsteady <strong>Flow</strong> Solutions<br />

Drag coefficient: C d a Strouhal number: St Wavelength: � Vortex speed: St �<br />

Gresho Gresho Berger Gresho Gresho<br />

Present et al. Clift Present et al. Roshko & Wille Present et al. Present et al.<br />

Re results [17] b et al. [5] c results [17] b [26] c [2] c results [17] b results [17] b<br />

50 d 1.38 1.81 1.41 0.139 0.14 0.122 0.12–0.13 6.9 6.5 0.96 0.93<br />

65 1.33 — 1.33 0.152 — 0.143 — 6.3 — 0.96 —<br />

100 1.26 1.76 1.24 0.171 0.18 0.167 0.16–0.17 5.5 5.2 0.94 0.93<br />

200 1.18 1.76 1.16 0.197 0.21 — 0.18–0.19 4.7 4.4 0.93 0.92<br />

400 1.18 1.78 1.12 0.22 0.22 0.205 0.2–0.21 4.2 4.4 0.92 0.96<br />

a Drag coefficient is averaged <strong>in</strong> the unsteady cases.<br />

b Computational study.<br />

c Experiment.<br />

d Unsteady solution (Fig. 7).<br />

(Figs. 5 and 6). The physical parameters <strong>of</strong> the separation tions, the flow field was numerically perturbed. Similarly,<br />

bubble are compared with previous experimental and com- <strong>in</strong> the present study a perturbation was needed <strong>in</strong> order<br />

putational studies <strong>in</strong> Table I and show good agreement to obta<strong>in</strong> vortex shedd<strong>in</strong>g at Re � 50; however, the higher<br />

for both Reynolds numbers.<br />

Reynolds number cases did not require any external per-<br />

When the Reynolds number is <strong>in</strong>creased to values <strong>in</strong> turbations. Consequently, Re � 50 is close to a critical<br />

the 35 � Re � 60 range, experiments observe asymmetry Reynolds number above which the solution becomes un-<br />

<strong>of</strong> the separation bubble and ‘‘wavy’’ behavior <strong>of</strong> the tail steady. It is possible that the steady state solution obta<strong>in</strong>ed<br />

<strong>of</strong> the wake. In the current computations it was found that for Re � 50 may eventually exhibit unstead<strong>in</strong>ess due to<br />

for Reynolds numbers <strong>in</strong> this range (<strong>in</strong> particular, Re � the buildup <strong>of</strong> truncation and mach<strong>in</strong>e errors <strong>in</strong> the solu-<br />

50) the type <strong>of</strong> wake behavior depends on whether the tion, but the computer time needed to arrive at this po<strong>in</strong>t<br />

<strong>in</strong>itial conditions are perturbed. For example, if no forced would be considerable.<br />

perturbations are imposed a steady separation bubble de- Accord<strong>in</strong>g to experimental results, <strong>in</strong>creas<strong>in</strong>g the Reynvelops<br />

downstream <strong>of</strong> the cyl<strong>in</strong>der (Fig. 7). The length <strong>of</strong> olds number beyond Re � 60 leads to the development<br />

the separation bubble <strong>in</strong> this case is L � 3 <strong>in</strong> agreement <strong>of</strong>aKármán vortex street forced by vortices which are<br />

with the steady state computation <strong>of</strong> Fornberg [13] at shed alternately with a dist<strong>in</strong>ct frequency from the top and<br />

Re � 50. If the solution shown <strong>in</strong> Fig. 7 is disturbed by bottom <strong>of</strong> the cyl<strong>in</strong>der (Coutanceau and Defaye [7]). In<br />

<strong>in</strong>troduc<strong>in</strong>g a small perturbation <strong>in</strong> the flow field, vortex the present study, the formation <strong>of</strong> the vortex street is<br />

shedd<strong>in</strong>g is <strong>in</strong>stigated. This behavior is demonstrated <strong>in</strong> depicted clearly <strong>in</strong> spanwise vorticity contours for Re �<br />

Fig. 8, where the cyl<strong>in</strong>der was moved vertically 0.001 nondi- 65, 100, 200, and 400 (Fig. 9). In Fig. 10, vertical velocity<br />

mensional units away from its orig<strong>in</strong>al position and then contours for Re � 400 are plotted, <strong>in</strong>dicat<strong>in</strong>g a very smooth<br />

moved back, perturb<strong>in</strong>g the flow field and generat<strong>in</strong>g a solution that is free <strong>of</strong> any detectable residual oscillations.<br />

Kármán vortex street. At Re � 50, Gresho et al. [17] also As the Reynolds number <strong>in</strong>creases, the frequency <strong>of</strong> the<br />

observed vortex shedd<strong>in</strong>g from the cyl<strong>in</strong>der with no shedd<strong>in</strong>g <strong>in</strong>creases (Table II) and the vortices become<br />

attached separation bubble, and the wake characteristics more concentrated. The patterns obta<strong>in</strong>ed for Re � 65<br />

measured <strong>in</strong> the unsteady solution <strong>of</strong> the current study and Re � 100 show remarkable similarity to the flow visualshow<br />

good agreement compared with Gresho et al.’s [17] izations <strong>of</strong> Freymuth et al. [14]. The time spectra and signa-<br />

results (Table II). In the computational studies <strong>of</strong> Jordan ture <strong>of</strong> the streamwise velocity at a po<strong>in</strong>t downstream <strong>of</strong><br />

and Fromm [20] and Braza et al. [4], no vortex shedd<strong>in</strong>g the cyl<strong>in</strong>der reveals the presence <strong>of</strong> a spike at the vortex<br />

or asymmetry <strong>of</strong> the separation bubble was observed for street Strouhal number (St � U� f/D : f � dimensional<br />

Reynolds numbers up to 1000. Braza et al. [4] expla<strong>in</strong>ed frequency); higher harmonics <strong>of</strong> the Strouhal number are<br />

this behavior by stat<strong>in</strong>g that the computational scheme also present (Fig. 11). Experiments predict that the onset<br />

was too ‘‘clean,’’ i.e., no external perturbations existed <strong>of</strong> three-dimensionality and turbulence will occur at Reyn-<br />

(as would appear <strong>in</strong> an experiment), therefore there was olds numbers below Re � 200 (Coutanceau and Defaye<br />

noth<strong>in</strong>g to trigger any asymmetry or unstead<strong>in</strong>ess <strong>of</strong> the [7]). Because <strong>of</strong> the two-dimensional nature <strong>of</strong> the current<br />

flow field. To <strong>in</strong>duce the vortex shedd<strong>in</strong>g <strong>in</strong> their simula- computations, turbulence and the effects <strong>of</strong> three-dimen-


VIRTUAL BOUNDARY METHOD 459<br />

FIG. 9. Spanwise vorticity; dotted and solid l<strong>in</strong>es denote negative and positive levels, respectively. (a) Re � 65; (b) Re � 100; (c) Re � 200;<br />

(d) Re � 400.<br />

sionality cannot be obta<strong>in</strong>ed, however, the vortices shed Tables II and III summarize the drag coefficient (Cd), from the cyl<strong>in</strong>der at Re � 200 and 400 exhibit some irregu- Strouhal number, the wave-length <strong>of</strong> the Kármán vortex<br />

larities associated with higher harmonics which subside as street (�), and the vortex speed (St �) observed for the<br />

the vortices are convected downstream, form<strong>in</strong>g a lam<strong>in</strong>ar unsteady solutions and they provide comparisons with pre-<br />

Kármán vortex street (Figs. 9c–d). vious computational and experimental results. The Strou-


460 SAIKI AND BIRINGEN<br />

FIG. 10. Re � 400 : normal velocity, dotted and solid l<strong>in</strong>es denote negative and positive levels, respectively.<br />

FIG. 11. Spectra and time signatures <strong>of</strong> streamwise velocity measured at (x d, y d) � 0.9445, �0.325); x d is the streamwise distance downstream<br />

from the cyl<strong>in</strong>der and y d is the normal distance from the symmetry axis <strong>of</strong> the cyl<strong>in</strong>der. (a) Re � 50 (unsteady solution, Fig. 7); (b) Re � 65; (c)<br />

Re � 100; (d) Re � 200; (e) Re � 400.


VIRTUAL BOUNDARY METHOD 461<br />

TABLE III the <strong>in</strong>flow boundary was too close to the cyl<strong>in</strong>der or if the<br />

Comparison <strong>of</strong> Quantitative Data with Previous<br />

Computational Results for Re � 100; Unsteady <strong>Flow</strong> Solutions<br />

Gresho Engelman Braza Joran &<br />

doma<strong>in</strong> was not wide enough, a higher Strouhal number<br />

was obta<strong>in</strong>ed. The distance between the <strong>in</strong>flow boundary<br />

and the cyl<strong>in</strong>der <strong>in</strong> the current study (Li � 4) is comparable<br />

to the doma<strong>in</strong> length used by Gresho et al. [17] and Engel-<br />

Present et al. & Jam<strong>in</strong>ia et al. Fromm Karniadakis man and Jam<strong>in</strong>ia [11], but shorter than those used by<br />

Re � 100 results [17] [11] [4] [20] et al. [21] Karniadakis and Triantafyllou [21], Braza et al. [4], and<br />

Cd St<br />

�<br />

St �<br />

1.26<br />

0.171<br />

5.5<br />

0.94<br />

1.76<br />

0.18<br />

5.2<br />

0.93<br />

1.411<br />

0.173<br />

5.32<br />

0.915<br />

1.28<br />

0.16<br />

—<br />

—<br />

1.28<br />

0.16<br />

—<br />

—<br />

—<br />

0.168<br />

—<br />

—<br />

Jordan and Fromm [20]. Accord<strong>in</strong>gly, as shown <strong>in</strong> Table<br />

III, the Strouhal number obta<strong>in</strong>ed <strong>in</strong> the current computations<br />

falls with<strong>in</strong> the range <strong>of</strong> Strouhal numbers determ<strong>in</strong>ed<br />

by the previous computational studies. We performed several<br />

test calculations with an expanded computational doma<strong>in</strong><br />

which revealed a slight drop <strong>in</strong> Strouhal number<br />

(from 0.171 to 0.168 for Re � 100) confirm<strong>in</strong>g that the<br />

hal numbers obta<strong>in</strong>ed from the present results are slightly proximity <strong>of</strong> the boundaries affects the vortex-shedd<strong>in</strong>g<br />

higher than the experimental results; however, they corre- frequency.<br />

spond better than the values obta<strong>in</strong>ed by the majority In the present study the drag coefficient was calculated<br />

<strong>of</strong> the other computational studies. The higher Strouhal <strong>in</strong> a manner similar to Goldste<strong>in</strong> et al. [16]; the drag was<br />

number can be attributed to the size <strong>of</strong> the computational found by consider<strong>in</strong>g the loss <strong>of</strong> fluid momentum <strong>in</strong> the<br />

doma<strong>in</strong>. Karniadakis and Triantafyllou [21] found that if doma<strong>in</strong>, i.e.,<br />

FIG. 12. Re � 200 : rotat<strong>in</strong>g cyl<strong>in</strong>der. Time evolution <strong>of</strong> streamfunction contours. (a) t � 1; (b) t � 1.5; (c) t � 2; (d) t � 3; (e) t � 4; (f) t �<br />

5; (g) t � 5.5; (h) t � 6.


462 SAIKI AND BIRINGEN<br />

FIG. 12—Cont<strong>in</strong>ued<br />

Cd � 2 �<br />

def<strong>in</strong>ed by the boundary po<strong>in</strong>ts <strong>of</strong> the cyl<strong>in</strong>der as � �<br />

�<br />

��<br />

U<br />

�1 �<br />

Ui U<br />

dy.<br />

Ui� (8) tan�1 ( yb/xb). The characteristics <strong>of</strong> the startup flow at this Reynolds<br />

Due to the <strong>in</strong>fluence <strong>of</strong> the boundary conditions on the<br />

computation, Ui was def<strong>in</strong>ed as the maximum mean veloc-<br />

ity measured at a distance D/2 upstream <strong>of</strong> the cyl<strong>in</strong>der.<br />

As Tables II and III reveal, good agreement is found for<br />

values <strong>of</strong> Cd, <strong>in</strong> comparison with experiments and previous<br />

computational work.<br />

number and rotation rate consist <strong>of</strong> a primary eddy evolv<strong>in</strong>g<br />

at the top <strong>of</strong> the cyl<strong>in</strong>der and a develop<strong>in</strong>g second eddy below<br />

the x-axis <strong>of</strong> symmetry (Figs. 12a–d). The second eddy<br />

moves upward (Figs. 12c–d) <strong>in</strong>duc<strong>in</strong>g two secondary vortices<br />

which merge to form a s<strong>in</strong>gle vortex at time, t � 6 (Figs.<br />

12e–h). The streamfunction contours obta<strong>in</strong>ed by the cur-<br />

rent study are <strong>in</strong> remarkable agreement with experimental<br />

Rotat<strong>in</strong>g <strong>Cyl<strong>in</strong>der</strong> <strong>in</strong> <strong>Uniform</strong> <strong>Flow</strong><br />

observations (Coutanceau and Ménard [8]) and previous<br />

computational results (Badr and Dennis [1]). Figure 13<br />

The startup flow over a cyl<strong>in</strong>der rotat<strong>in</strong>g counterclock- demonstrates excellent comparison between pr<strong>of</strong>iles <strong>of</strong><br />

wise <strong>in</strong> uniform flow was computed for Re � 200. The streamwise and normal velocity along the x-axis beh<strong>in</strong>d the<br />

rotation rate <strong>of</strong> the cyl<strong>in</strong>der was � � 1, result<strong>in</strong>g <strong>in</strong> a cyl<strong>in</strong>der obta<strong>in</strong>ed from the current computation and experi-<br />

tangential velocity <strong>of</strong> one half the freestream velocity mental measurements (Coutanceau and Ménard [8]).<br />

(vt � 0.5). The motion <strong>of</strong> the cyl<strong>in</strong>der was <strong>in</strong>troduced<br />

by sett<strong>in</strong>g the components <strong>of</strong> v <strong>in</strong> Eq. (6) to the proper<br />

Oscillat<strong>in</strong>g <strong>Cyl<strong>in</strong>der</strong> <strong>in</strong> <strong>Uniform</strong> <strong>Flow</strong><br />

streamwise and normal velocities aris<strong>in</strong>g from the rotation This computation was performed at Re � 200 with the<br />

<strong>of</strong> the cyl<strong>in</strong>der, i.e., v � (�vt s<strong>in</strong> �, vt cos �), where � is<br />

cyl<strong>in</strong>der oscillat<strong>in</strong>g parallel to the free-stream velocity at


VIRTUAL BOUNDARY METHOD 463<br />

FIG. 13. Re � 200 : rotat<strong>in</strong>g cyl<strong>in</strong>der. Time evolution <strong>of</strong> the velocity pr<strong>of</strong>iles along the x-axis <strong>of</strong> symmetry. (a) streamwise velocity; (b) normal<br />

velocity. Current computational results: ----------. t � 1; –––, t � 1.5; –-–-–-, t � 2; –---–---, t � 3; ——, t � 4. Experimental results [8]: �, t � 1;<br />

*, t � 1.5; �, t � 2; �, t � 3; �, t � 4.<br />

a frequency, fc � 1.88 St, i.e., 1.88 times the Strouhal fre- the two clockwise vortices mov<strong>in</strong>g downstream alongside<br />

quency for the stationary cyl<strong>in</strong>der. The amplitude <strong>of</strong> the the counterclockwise eddy. These results are <strong>in</strong> excellent<br />

oscillation resulted <strong>in</strong> a streamwise displacement <strong>of</strong> the agreement with experimental observations (Griff<strong>in</strong> and<br />

cyl<strong>in</strong>der <strong>of</strong> �0.24, and the cyl<strong>in</strong>der motion was prescribed<br />

by sett<strong>in</strong>g the horizontal velocities on the boundary po<strong>in</strong>ts<br />

Ramberg [18]; Ongoren and Rockwell [22]).<br />

to ub � Ac cos(2�fct). The computations started with the<br />

cyl<strong>in</strong>der stationary and oscillations were imposed once the<br />

CONCLUSION<br />

solution reached a quasi-steady state. In this study, we applied a virtual boundary method to<br />

For the parameters considered <strong>in</strong> the current study, several steady/unsteady flow problems. The method modthe<br />

vortex shedd<strong>in</strong>g pattern <strong>of</strong> the stationary case at els a no-slip boundary by an external forc<strong>in</strong>g function<br />

Re � 200 (Fig. 9c) is modified by the oscillation <strong>of</strong> the added to the momentum equations. The computational<br />

cyl<strong>in</strong>der. This clearly depicted <strong>in</strong> the time evolution <strong>of</strong> results for both stationary and mov<strong>in</strong>g cyl<strong>in</strong>ders <strong>in</strong> uniform<br />

spanwise vorticity contours over two oscillation periods flow compare favorably with both experimental and previ-<br />

<strong>of</strong> the cyl<strong>in</strong>der (Fig. 14). Dur<strong>in</strong>g this time period an ous computational studies and lend further evidence to the<br />

antisymmetrical mode A � III (Ongoren and Rockwell applicability <strong>of</strong> the virtual boundary technique for steady<br />

[22]) appears consist<strong>in</strong>g <strong>of</strong> two clockwise vortices shed and unsteady flow problems. The oscillations caused by<br />

from the top <strong>of</strong> the cyl<strong>in</strong>der and the evolution <strong>of</strong> a the virtual boundary method when used with a spectral<br />

s<strong>in</strong>gle counterclockwise eddy from the bottom. These discretization method were attenuated by the application<br />

vortices then form a vortex street with the weaker <strong>of</strong><br />

<strong>of</strong> high-order f<strong>in</strong>ite differences.


464 SAIKI AND BIRINGEN<br />

FIG. 14. Re � 200 : oscillat<strong>in</strong>g cyl<strong>in</strong>der. Time evolution <strong>of</strong> spanwise vorticity contours; dotted and solid l<strong>in</strong>es denote negative and positive<br />

contours, respectively. (a) t � T/4; (b) t � T/2; (c) t � 3T/4; (d) t � T; (e) t � 5T/4; (f) t � 3T/2; (g) t � 7T/4; (h) t � 2T (T is the oscillation<br />

period <strong>of</strong> the cyl<strong>in</strong>der).


VIRTUAL BOUNDARY METHOD 465<br />

ACKNOWLEDGMENTS<br />

Concentric <strong>Cyl<strong>in</strong>der</strong>s with an External Force Field,’’ <strong>in</strong> Proceed<strong>in</strong>gs,<br />

11th AIAA Computational Fluid Dynamices Conference, Orlando,<br />

This work was supported by the Navy under Grant DOD N00014-94- Florida, July 1993.<br />

1-0923. The authors thank James A. Fe<strong>in</strong> for his cont<strong>in</strong>ued <strong>in</strong>terest <strong>in</strong> 16. D. Goldste<strong>in</strong>, R. Handler, and L. Sirovich, J. Comput. Phys. 105,<br />

this work.<br />

354 (1993).<br />

17. P. M. Gresho, R. Chan, C. Upson, and R. Lee, Int. J. Numer. Methods<br />

REFERENCES<br />

Fluids 4, 619 (1984).<br />

18. O. M. Griff<strong>in</strong> and S. E. Ramberg, J. Fluid Mech. 75(2), 31 (1976).<br />

1. H. M. Badr and S. C. R. Dennis, J. Fluid Mech. 158, 447 (1985). 19. A. Huser and S. Bir<strong>in</strong>gen, Int. J. Numer. Methods Fluids 14, 1087<br />

2. E. Berger and R. Wille, Annu. Rev. Fluid Mech. 4, 313 (1972).<br />

(1992).<br />

3. R. P. Beyer, J. Comput. Phys. 98, 145 (1992).<br />

20. S. K. Jordan and J. E. Fromm, Phys. Fluids 15, 371 (1972).<br />

4. M. Braza, P. Chassa<strong>in</strong>g, and H. Ha M<strong>in</strong>h, J. Fluid Mech. 165, 79 (1986).<br />

21. G. E. Karniadakis and G. S. Triantafyllou, J. Fluid Mech. 238, 1 (1992).<br />

5. R. Clift, J. R. Grace, and M. E. Weber, Bubbles, Drops, and Particles<br />

(Academic Press, New York, 1978), p. 154.<br />

22. A. Ongoren and D. Rockwell, J. Fluid Mech. 191, 225 (1988).<br />

23. C. S. Pesk<strong>in</strong>, J. Comput. Phys. 25, 252 (1977).<br />

6. M. Coutanceau and R. Bouard, J. Fluid Mech. 79(2), 231 (1977). 24. C. S. Pesk<strong>in</strong>, Annu. Rev. Fluid Mech. 14, 235 (1982).<br />

7. M. Coutanceau and J. R. Defaye, Appl. Mech. Rev. 44, 255 (1991).<br />

25. A. Roshko, NACA Technical Note 2913, 1953 (unpublished).<br />

8. M. Coutanceau and C. Ménard, J. Fluid Mech. 158, 399 (1985).<br />

9. G. Danabasoglu, Ph.D. thesis, University <strong>of</strong> Colorado, 1992 (unpublished).<br />

10. G. Danabasoglu, S. Bir<strong>in</strong>gen, and C. L. Streett, Phys. Fluids A 3,<br />

2138 (1991).<br />

26. E. M. Saiki, S. Bir<strong>in</strong>gen, G. Danabasoglu, and C. L. Streett, J. Fluid<br />

Mech. 253, 485 (1993).<br />

27. E. M. Saiki and S. Bir<strong>in</strong>gen, ‘‘<strong>Numerical</strong> <strong>Simulation</strong> <strong>of</strong> Particle Effects<br />

on Boundary Layer <strong>Flow</strong>,’’ <strong>in</strong> Transition, Turbulence, and Combustion,<br />

Volume I: Transition edited by M. Y. Hussa<strong>in</strong>i, T. B. Gatski,<br />

and T. L. Jackson (Kluwer Academic, Dordrecht, 1994, p. 267.<br />

11. M. S. Engelman and M. Jam<strong>in</strong>ia, Int. J. Numer. Methods Fluids 11,<br />

985 (1990).<br />

28. L. Sirovich, Phys. Fluids 10, 24 (1967).<br />

29. L. Sirovich, Phys. Fluids 11, 1424 (1968).<br />

12. A. L. Fogelson and C. S. Pesk<strong>in</strong>, J. Comput. Phys. 79, 50 (1988).<br />

30. C. L. Streett and M. Y. Hussa<strong>in</strong>i, ICASE Report No. 86-59, 1986 (un-<br />

13. B. Fornberg, J. Fluid Mech. 98, 819 (1980). published).<br />

14. P. Freymuth, F. F<strong>in</strong>aish, and W. Bank, Phys. Fluids 29, 1321 (1986). 31. C. L. Streett and M. G. Macaraeg, Appl. Numer. Math. 6, 123 (1989).<br />

15. D. Goldste<strong>in</strong>, T. Adachi, and H. Sakata, ‘‘Model<strong>in</strong>g a <strong>Flow</strong> between<br />

32. C. Tu and C. S. Pesk<strong>in</strong>, SIAM J. Sci. Stat. Comput. 13, 1361 (1992).

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

Saved successfully!

Ooh no, something went wrong!