Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

A numerical investigation of wall effects up to high blockage ratios on two-

dimensional flow past a confined circular cylinder


Mehmet Sahin and Robert G. Owens

Citation: Phys. Fluids 16, 1305 (2004); doi: 10.1063/1.1668285


View online: http://dx.doi.org/10.1063/1.1668285
View Table of Contents: http://pof.aip.org/resource/1/PHFLE6/v16/i5
Published by the AIP Publishing LLC.

Additional information on Phys. Fluids


Journal Homepage: http://pof.aip.org/
Journal Information: http://pof.aip.org/about/about_the_journal
Top downloads: http://pof.aip.org/features/most_downloaded
Information for Authors: http://pof.aip.org/authors

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
PHYSICS OF FLUIDS VOLUME 16, NUMBER 5 MAY 2004

A numerical investigation of wall effects up to high blockage ratios


on two-dimensional flow past a confined circular cylinder
Mehmet Sahin and Robert G. Owensa)
LMF-ISE-FSTI, Ecole Polytechnique Fédérale de Lausanne, CH 1015 Lausanne, Switzerland
共Received 25 September 2003; accepted 20 January 2004; published online 2 April 2004兲
A finite volume method based on a velocity-only formulation is used to solve the flow field around
a confined circular cylinder in a channel in order to investigate lateral wall proximity effects on
stability, Strouhal number, hydrodynamic forces and wake structure behind the cylinder for a wide
range of blockage ratios (0.1⬍ ␤ ⭐0.9) and Reynolds numbers (0⬍Re⭐280). For blockage ratios
less than approximately 0.85 a first critical Reynolds number is identified at which a supercritical
Hopf bifurcation of the symmetric solution occurs. For blockage ratios greater than about 0.687 and
at Reynolds numbers exceeding the first critical Reynolds number a second curve of neutral stability
is seen, representing a pitchfork bifurcation of the steady symmetric solution to one of two possible
steady asymmetric solutions. Either side of the neutral stability curve for the pitchfork bifurcation
our linear stability analysis and direct numerical simulations demonstrate that although the flow is
linearly stable it is unstable to finite two-dimensional perturbations. At blockage ratios larger than
about 0.82 the steady asymmetric solutions also become unstable through a Hopf bifurcation. In
contrast with the first Hopf bifurcation of the symmetric solution at lower Reynolds numbers
numerical calculations of the lift coefficient reveal that the oscillations are no longer symmetric in
the rising and falling parts of each cycle. Very strong vortices shed from the cylinder and the wall
cause drastic increases in the amplitudes of the lift and drag coefficients. A co-dimension 2 point
where pitchfork and Hopf bifurcations occur simultaneously has been located in parameter space.
Altogether, four distinct regions in the parameter space ( ␤ ,Re)苸(0,0.9 兴 ⫻(0,280兴 have been
identified, each corresponding to a different class of flow: 共i兲 Steady symmetric flow, 共ii兲 symmetric
vortex shedding, 共iii兲 steady asymmetric flow, and 共iv兲 asymmetric vortex shedding, where a
periodic-in-time flow is classed as symmetric or asymmetric depending on whether the time-average
over one cycle of the lift coefficient is zero or not. Numerical solutions are computed on meshes
having up to 1.8 million degrees of freedom. Extensive comparisons are made with the results
available in the literature. © 2004 American Institute of Physics. 关DOI: 10.1063/1.1668285兴

I. INTRODUCTION we have in mind is depicted in Fig. 1. In this figure an


infinitely long cylinder of diameter D is placed symmetri-
It is no exaggeration to say that an enormous 共and still
cally between parallel lateral walls a distance H apart. The
rapidly growing兲 corpus of literature on the subject of bluff
parameter ␤ ⬅D/H is usually termed the blockage ratio. In
body wakes has developed since the pioneering work of von
Kármán early last century. This fact is an attestation to both stark contrast to the wealth of insight and commentary avail-
the difficulty in understanding and adequately describing the able on vortex dynamics in the wake of an unbounded cyl-
flow bifurcations that occur at various values of the Reynolds inder we find ourselves with only a handful of papers offer-
number in viscous flows and the interest in doing so. Flows ing a serious treatment of the blockage ratio effects present
having particularly simple setups such as those past a sphere in the confined cylinder problem. This paucity of scientific
or cylinder have succeeded in drawing experimentalists, literature should not be interpreted as implying that the prob-
theoreticians and computational fluid dynamicists into the lem is an unimportant one, however. On the contrary, even
fray that has gone on through the decades and only very for unbounded flow past a cylinder the 共infinite兲 flow domain
recently are consensuses emerging. has to be replaced with 共or mapped onto兲 a finite one, thus
Details of recent theoretical, experimental and computa- introducing numerical or experimental blockage effects that
tional developments for unbounded flow past a cylinder may
may have considerable influence over the determined values
be found in the review paper of Williamson,1 where particu-
of the flow parameters.2,3 Many of the blockage ratio effects
lar attention is paid to the vortex dynamics in the cylinder
wake. Our interest in this paper is a careful analysis of lateral described in the literature are more or less evident:
wall effects on viscous flow past a confined cylinder. What 共1兲 In the steady flow regime, bringing the walls closer to
the cylinder results in the appearance of the twin vortices
a兲
Author to whom correspondence should be addressed. Electronic mail: in the cylinder wake at higher Reynolds numbers.
robert.owens@epfl.ch 共2兲 At any given modest (ⱗ50) Reynolds number and for

1070-6631/2004/16(5)/1305/16/$22.00 1305 © 2004 American Institute of Physics

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
1306 Phys. Fluids, Vol. 16, No. 5, May 2004 M. Sahin and R. G. Owens

number of 4.8⫻104 . Lei et al.7 also noted only slight fluc-


tuations in a Strouhal number computed from the free-stream
velocity for a similar range of gap to diameter ratios. How-
ever, for gaps less than 0.3 cylinder diameters8 or 0.2–0.3
diameters 共depending on the boundary layer thickness7兲, vor-
tex shedding was suppressed. Differences in the quantifica-
tion of the vortex shedding suppression gap ratio were due
possibly to differences in the boundary layer thicknesses
generated by the experimentalists and also to the manner in
FIG. 1. Schematic of a cylinder placed symmetrically in a plane channel. which the critical gap ratio was identified: Bearman and
The cylinder diameter is D and the channel height H. Zdravkovich8 using a spectral analysis of hot-wire signals in
the cylinder wake whereas the method of Lei et al.7 was
based on observation of the spectrum of the lift coefficient.
␤ ⭐0.2 the length of the closed vortex bubble decreases Suppression of vortices for a sufficiently small gap ratio was
with wall proximity, while remaining a linear function of also confirmed by Zovatto and Pedrizzetti,9 who used a finite
Re. 2– 4 element method based on a vorticity-streamfunction formu-
共3兲 For increasing blockage ratios ␤ up to 0.5 the steady lation to analyze flow around a cylinder positioned eccentri-
two-dimensional base flow is stabilized with respect to cally between two lateral walls. For very small gap ratios
infinitesimal perturbations due to constraint by the con- Zovatto and Pedrizzetti9 found a recirculating bubble on the
fining walls of the separating shear layer that exists be- wall downstream of the cylinder. A separation bubble on the
tween the cylinder wake and the wall boundary layer wall had also been seen earlier by Bearman and
vorticity.3,4 Zdravkovich8 for gap ratios smaller than the critical value for
共4兲 Once the critical Reynolds number for the primary insta- vortex suppression. We will return to some of these flow
bility has been exceeded the frequency with which peri- phenomena in our discussion of our numerical results in Sec.
odic two-dimensional vortex shedding takes place at a IV for large blockage ratios.
given Reynolds number is an increasing function of
The motivation for the present study is twofold. First,
␤.2,4 – 6 共Note that the spurious result obtained by Stansby
the rich fluid dynamics in the wake and near the lateral walls
and Slaouti6 for ␤ ⫽0.5 is thought to be due to neglect of
deserves to be investigated with greater numerical accuracy
the boundary layers in their numerical simulation using
than has been possible with the computational resources
random vortex methods.兲
available to other researchers at the time at which they pre-
共5兲 At Re⫽O(100) both the mean drag coefficient C d and
pared their manuscripts. Computations on meshes allowing
the separation angle of the vortex bubble increase as the
for only tens of thousands of degrees of freedom have been
walls approach the cylinder.2,3,5,6
typical 共for example, Refs. 4 and 10兲. In the present study a
In addition to the obvious interest of wall blockage effects novel finite volume method11–13 is used in a parallel imple-
and as observed by Chen et al.,4 the choice of a bounded mentation, permitting up to 1.8 million degrees of freedom
domain allows a more definitive specification of the flow and thus a higher resolution of the wake and boundary layer
共both numerically and experimentally兲 than is possible in the structures. Second, the only previous numerical linear stabil-
unbounded case, whilst conserving the essential features of ity analysis of flow past a confined cylinder available to us4
the latter. went no further than a blockage ratio of ␤ ⫽0.7. From the
A problem bearing some similarities to that of flow past results of this publication the trend seemed to be one of
a confined cylinder is that of flow around a cylinder placed at decreasing linear stability of the two-dimensional flow for
various heights above a plane boundary. A recent literature ␤ ⬎0.5. Stability was always lost over the range of blockage
survey of experimental investigations into this problem may ratios considered through a symmetry-breaking supercritical
be found in the paper of Lei et al.7 These studies have sought Hopf bifurcation. We wish in this paper to investigate the
to address the issue of how forces on the cylinder and vortex effect on the critical Reynolds number of choosing ␤ ⬎0.7
shedding frequency depend on the ratio g/D of the gap be- and to identify the nature of the flow instabilities by means
tween the cylinder and the wall, g, and the cylinder diameter, of an Arnoldi method.
D. They have also been concerned with understanding the The outline of the present paper is as follows: In Sec. II
effect on these quantities of the boundary layer thickness and we describe the problem to be solved and furnish the reader
the velocity gradient. Most of the experiments have been with a brief description of the numerical method used to
conducted at Reynolds numbers in the sub-critical regime analyze flow past a confined cylinder at Reynolds numbers
关 Re⫽O(1⫻104 ) 兴 in which the boundary layer is still lami- up to 280. Section III is dedicated to validation of our nu-
nar. Lei et al.7 found that the drag coefficient C d increased merical scheme for the classical problem of unbounded two-
with increasing gap ratio because of the reduction in the base dimensional flow past a circular cylinder. Extensive compari-
pressure. The same trend in base pressure dependence had son with other results in the literature is made. In particular
been observed by Bearman and Zdravkovich.8 The latter au- we find excellent agreement with previously obtained values
thors further found that the Strouhal number for g/Dⲏ0.3 for the drag coefficients, first critical Reynolds numbers
was more or less constant in their experiments at a Reynolds Re crit1 and the corresponding critical Strouhal numbers. In

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
Phys. Fluids, Vol. 16, No. 5, May 2004 A numerical investigation of wall effects 1307

Sec. IV we are concerned with a detailed description of wake ⳵ 2u 1 ⳵u2


dynamics and interactions of the wake and wall boundary Outflow: 2 ⫽0, ⫽0. 共10兲
⳵x1 ⳵x1
layers for blockage ratios up to 0.9. For blockage ratios be-
low approximately 0.85 the locus of a supercritical Hopf Some care needs to be taken with how the second normal
bifurcation may be traced out in parameter space. At higher derivative outflow condition is imposed, due to possible lin-
Reynolds numbers and for blockage ratios sufficiently large ear dependence in the discrete equation set of this condition
there is a pitchfork bifurcation of the steady symmetric state with the discrete form of the continuity equation. More pre-
to one of two asymmetric steady states. Either side of the cisely stated, the second normal derivative outflow condition
curve of neutral stability for the pitchfork bifurcation the will be automatically satisfied at x 2 ⫽0 upon imposition of
steady solutions are linearly stable but appear on the basis of the continuity equation 共2兲 within each finite volume. For all
direct numerical simulations to be unstable to finite two- the results presented in Sec. III the dimensionless upstream
dimensional perturbations. For yet larger Reynolds numbers and downstream channel lengths were set equal to 400 D. In
and ␤ ⲏ0.82 a Hopf bifurcation of the asymmetric state oc- Sec. IV these lengths were chosen to both be equal to 40 D.
curs. The oscillations are now quite different from those as- The choice of outflow boundary conditions 共6兲 and 共10兲 was
sociated with the first symmetry-breaking instability, the am- motivated by the fact that our numerical method uses a
plitude of the drag and lift coefficients being much stronger velocity-only formulation so that the usual traction-free con-
and the oscillations are now asymmetric in time. Finally, we ditions could not easily be implemented. The free boundary
draw some conclusions. layer type of conditions 共6兲 and 共10兲 were used successfully
by Kourta et al.14 in finite volume simulations of a two-
II. MATHEMATICAL PROBLEM AND NUMERICAL dimensional plane mixing layer. Although Jin and Braza15
SCHEME later developed a nonreflecting outlet condition that greatly
reduced feedback noise when compared with the outlet con-
An infinitely long cylinder of diameter D is placed mid-
dition of Kourta et al.,14 the outlet length used for the calcu-
way between two parallel planes which are a distance H
lations in the present paper are considered sufficiently great
apart, as shown in Fig. 1. Let us denote by U max the maxi-
that the difference between the influence of the one set of
mum inlet fluid speed. The incompressible unsteady Navier–
exit conditions and the other on drag, linear stability and
Stokes equations may be written in dimensionless form as
Strouhal number would be negligible. The more complicated
⳵u 1 2 exit conditions of Jin and Braza15 are therefore not imple-
⫹ 共 u•ⵜ 兲 u⫽⫺ⵜ p⫹ ⵜ u, 共1兲
⳵t Re mented.
Let n denote a unit outward pointing normal vector to
ⵜ•u⫽0, 共2兲 the boundary ⳵⍀ of a finite volume ⍀. Then integration of
where, in the usual notation, u⫽(u 1 ,u 2 ) denotes the velocity 共2兲 over ⍀ and taking the vector product of 共1兲 with n, fol-
field, p the pressure and Re is a Reynolds number. In the lowed by integration around ⳵⍀ leads, respectively, to


present work the Reynolds number is defined as Re
⫽U maxD/v where v is the kinematic viscosity. In the presen- n"u ds⫽0 共11兲
tation of results in Secs. III and IV for those flows exhibiting ⳵⍀

periodic vortex shedding, a Strouhal number St is defined by


and
St⫽D/(TU max), where T is the period of vortex shedding.
We denote by (x,t)⫽((x 1 ,x 2 ),t) a generic point in space
and time. 冖 冋 ⳵⳵
⳵⍀
n⫻
u
t
⫹ 共 ⵜ⫻u兲 ⫻u⫹
1
Re
ⵜ⫻ 共 ⵜ⫻u兲 ds⫽0. 册
In Sec. III we approximate the unbounded cylinder ge- 共12兲
ometry by choosing ␤ ⫽0.01 and the following boundary
conditions: In our numerical scheme the continuity equation 共11兲 is sat-
isfied within each finite volume while 共12兲 is applied to each
Cylinder surface: u⫽ 共 0,0 兲 , 共3兲 finite volume except the finite volumes next to the wall.
Lateral walls: u⫽ 共 1,0 兲 , 共4兲 Therefore, vorticity creation is allowed within these finite
volumes in order to satisfy the no-slip boundary conditions.
Inflow: u⫽ 共 1,0 兲 , 共5兲 Equations 共11兲 and 共12兲 with no-slip boundary conditions are
⳵ 2u 1 ⳵u2 enough to solve the problem in a simply connected domain
Outflow: 2 ⫽0, ⫽0. 共6兲 共such as that found in the lid-driven cavity problem, for
⳵x1 ⳵x1
example12兲. However, if the domain is not simply connected
For the confined cylinder problem 共see Sec. IV兲 Eqs. 共1兲 and there is a need for additional equations. This is because there
共2兲 are solved subject to the following boundary conditions is a potential problem in our velocity-only formulation with
on the components of velocity: multi-valuedness of the pressure field, even though the pres-
sure does not appear explicitly as a dependent variable in our
Cylinder surface: u⫽ 共 0,0 兲 , 共7兲
formulation. To rectify this a Kutta-type condition

冖 冋 ⳵⳵ 册
Lateral walls: u⫽ 共 0,0 兲 , 共8兲
u 1
n⫻ ⫹ 共 ⵜ⫻u兲 ⫻u⫹ ⵜ⫻ 共 ⵜ⫻u兲 ds⫽0, 共13兲
Inflow: u⫽ 共 1⫺x 22 ,0兲 , 共9兲 ⌫ t Re

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
1308 Phys. Fluids, Vol. 16, No. 5, May 2004 M. Sahin and R. G. Owens

TABLE I. Values of grid parameters i max , k max , k wall , and N.

M1 M2 M3

␤ i max k max k wall N i max k max k wall N i max k max k wall N

0.01 181 301 137 89 336 361 601 273 355 312 721 1201 545 1 417 184
0.1 181 441 77 116 536 361 881 153 462 512 721 1761 305 1 842 784
0.2 181 421 57 109 336 361 841 113 433 712 721 1681 225 1 727 584
0.3 181 411 47 105 736 361 821 93 419 312 721 1641 185 1 669 984
0.5 181 401 37 102 136 361 801 73 404 912 721 1601 145 1 612 384
0.7 181 391 27 98 536 361 781 53 390 512 721 1561 105 1 554 784
0.9 181 381 17 94 936 361 761 33 376 112 721 1521 65 1 497 184

is imposed around the closed path ⌫ formed from the union where C⫽(A⫺␭M) ⫺1 M and ␮ ⫽( ␴ ⫺␭) ⫺1 . Application of
of the outer edges of the finite volumes on the cylinder sur- the Arnoldi method results in the construction of an upper
face. The condition 共13兲 guarantees that Hessenberg matrix whose eigenvalues are approximations to
a subset of the eigenvalues ␮ of C. From the properties of
冖 ⌫
n⫻ⵜpds⫽k关 p 兴 ⫽0, 共14兲 Arnoldi’s method and in the absence of a shift ␭,18 best reso-
lution of the ␴-spectrum is expected to be near the origin.
The coefficient matrix A in 共16兲 is almost identical 共by
where k is a unit vector normal to the plane of the flow, and
construction兲 to that which arises in the computations of the
关 p 兴 denotes the jump in the pressure on passing once around
steady base flow using Newton’s method. Solutions to all the
⌫. Since 共12兲 is satisfied in every interior finite volume, sat-
discrete algebraic equations that arise in the steady, unsteady
isfaction of 共13兲 ensures that p is single-valued at every in-
or eigenvalue problems of this paper have been obtained by
terior finite volume vertex. The pressure can be obtained by
implementing the MUltifrontal Massively Parallel Solver
integrating the two components of the pressure gradient ap-
共MUMPS兲 of Amestoy et al.19,20 The multifrontal method
pearing in the equations of linear momentum in a manner
used is a direct method based on LU decomposition for the
analogous to that used in finding a streamfunction from a
solution of sparse systems of linear equations with optimum
given velocity. The values of p on the domain boundaries,
fill in. The algorithms employed by MUMPS use a dynamic
when required, are determined by first computing ⳵ p/ ⳵ n
distributed task scheduling technique that permits numerical
from 共1兲.
pivoting and the transfer of computational tasks to lightly
A fully implicit second-order cell-vertex finite volume
loaded processors. The calculations have been performed on
method based on a velocity-only formulation is used for the
an SGI Origin 3800 parallel machine with 124 processors
discretization of 共1兲 and 共2兲. Discretization of the integrals
and on a Linux cluster with 22 processors.
appearing in 共11兲 and 共12兲 is effected by using the mid-point
Three different finite volume grids (M 1 – M 3) have
rule on cell faces. Full details of the method are supplied in
been used for each value of the blockage ratio considered in
two recent papers by the present authors11,12 and, in the in-
this paper. Each of the meshes has been generated algebra-
terests of brevity, will not be reproduced here. For the time-
ically and then smoothed by solving elliptic partial differen-
dependent computations presented in Secs. III and IV we
tial equations for the spatial variables x 1 and x 2 where de-
discretize in time using an Euler implicit method and for
rivatives are with respect to mapped variables in a space in
computing steady-state base flows a Newton method is em-
which the mesh appears rectangular.21 For the present prob-
ployed.
lem the physical grid is ‘‘cut’’ along the line x 2 ⫽0 from the
A major part of the present paper is concerned with the
rear stagnation point to the outlet before being mapped. The
linear stability of two-dimensional flow at various different
method of Steger and Sorenson21 allows both grid cell sizes
blockage ratios. Consider the perturbed flow
and grid cell skewness to be controlled at the inner and outer
u共 x,t 兲 ⫽U共 x兲 ⫹v共 x兲 exp共 ␴ t 兲 , 共15兲 boundaries. Meshes M 1 to M 3 are characterized by i max :
The number of nodes on the surface of the cylinder, k max :
where U(x) is the 共numerically determined兲 steady base flow The number of nodes along the line x 2 ⫽0 from the rear
at a given Reynolds number. Then discretizing the dimen- stagnation point on the cylinder to the outflow boundary and
sionless Navier–Stokes equations as described above leads k wall : The number of nodes in the gap between the cylinder
to an algebraic system of equations and a lateral wall. The values of i max , k max , k wall , and N 共the
Ax⫽ ␴ Mx, 共16兲 number of degrees of freedom兲 for the three meshes are sup-
plied in Table I for different blockage ratios.
for the nodal values of the perturbation velocity v. The ma-
trices A and M in 共16兲 are block quad-diagonal and block III. FLOW PAST AN UNBOUNDED CIRCULAR
bi-diagonal, respectively. The GEVP 共16兲 may be solved by CYLINDER
applying Arnoldi’s method16,17 to the equivalent system
Flow around an unbounded circular cylinder is a classi-
Cx⫽ ␮ x, 共17兲 cal benchmark problem for which a large number of numeri-

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
Phys. Fluids, Vol. 16, No. 5, May 2004 A numerical investigation of wall effects 1309

TABLE II. Unbounded flow past a cylinder. Comparison of critical Reynolds numbers computed on M1–M3
with others in the literature.

Ding and Noack and


M1 M2 M3 Extrapolated Jackson Kawahara Eckelmann Chen et al.

Re crit1 47.08 46.82 46.76 46.74 46.184 46.389 50 47.9


St crit1 0.1163 0.1166 0.1167 0.1167 0.138 04 0.126 19 0.132 0.138

cal and experimental results exists. In this problem, and in compute the leading eigenvalue with ␭⫽0.50⫾0.00i while
approximation to the case of an unbounded flow domain, a with no shift a Krylov subspace dimension larger than 1000
circular cylinder of diameter D⫽1.00 is placed symmetri- may be required.
cally in a channel with blockage ratio ␤ ⫽0.01. For the nu- The computed eigenspectrum on mesh M 3 at the critical
merical linear stability analysis the three meshes M 1 to M 3 Reynolds number is given in Fig. 2. Although we present the
were used, with i max , k max , and k wall as given in Table I. first 250 computed eigenvalues, calculations with higher
However, for unsteady time-dependent simulations we were Krylov subspace dimensions showed that only the leading
only able to afford to use M1 and M2, the unsteady calcula- eigenvalues and the eigenvalues around the origin were
tions on M3 proving to be prohibitively expensive. On the properly converged. As may be seen, the most dangerous
boundaries of the computational domain the conditions 共3兲– eigenvalue pair is well separated from the rest of the spec-
共6兲 were imposed. trum, unlike the eigenspectrum for the two-dimensional lid-
The linear stability analysis predictions of the critical driven cavity problem, for example.13 This is likely to be the
Reynolds and Strouhal numbers corresponding to the onset reason for well-developed periodic flow observed far beyond
of the first flow instability are supplied in Table II, as com- the critical Reynolds number. Our critical Strouhal number
puted on meshes M 1 – M 3. Also shown are the values of 共0.1167兲 compares very well with the Strouhal number St
these quantities when extrapolated to zero mesh size. The ⫽0.1179 computed at the same Reynolds number 共46.74兲
extrapolated critical Reynolds number is found to be Re crit1 from a curve fit of the two-dimensional experimental data of
⫽46.74 with a corresponding Strouhal number of St crit1
Williamson.23 In addition, our critical Strouhal number
⫽0.1167. These values are compared with others in the lit-
agrees quite well with the Strouhal number (St⫽0.118 34)
erature in the same table. Although we find good agreement
of the direct numerical simulation of Posdziech and
for the critical Reynolds number with the result of Ding and
Grundmann,24 even though their critical Strouhal number
Kawahara22 (Re crit1 ⫽46.389), and Jackson10 (Re crit1
and ours were computed at two slightly different Reynolds
⫽46.184), the critical Strouhal number manifests wider scat-
numbers (Re⫽47.50 and Re⫽46.74, respectively兲.
ter in the cited references. Issues such as the blockage ratios
chosen, distances from the cylinder of the upstream and In Fig. 3 we present comparisons of the Strouhal number
downstream boundaries, boundary conditions, mesh resolu- versus Reynolds number and in Fig. 4 comparisons of the
tion and number of eigenvalues determined may be amongst drag coefficient C d ⫽F x /0.5U max
2
D versus Reynolds number,
the reasons for discrepancies in the numerical results. In ad- in further verification of our numerical scheme. Our Strouhal
dition to our mesh convergence study, we present a conver- numbers are seen to be in very good agreement with those
gence study of the leading eigenvalues on mesh M 1 with the from the experimental work of Williamson23 for Reynolds
Krylov subspace dimension m and shift parameter ␭ in Table numbers up to 200. Beyond this point the flow becomes
III in order to show that our leading eigenvalue is essentially three-dimensional and we do not expect to have agreement
independent of both m and ␭ for sufficiently large values of with the experimental results. Good agreement for St and Cd
these two parameters. Although the leading eigenvalue con- with results from the two-dimensional numerical simulations
verges very rapidly with a suitably chosen complex shift of Henderson25 and Posdziech and Grundmann24 may also be
around the leading eigenvalue, it requires complex arith- seen from Figs. 3 and 4. Our computed lift coefficients C l at
metic. A real shift also dramatically improves the conver- Reynolds numbers of 100 and 200 are ⫾0.3333 and
gence of the leading eigenvalue while avoiding complex ⫾0.6861 and these are in satisfactory agreement with Pos-
arithmetic which significantly increases the memory require- dziech and Grundmann’s values of ⫾0.321 04 and
ments during LU factorization. Our calculations show that a ⫾0.673 15, respectively. Although both Henderson25 and
Krylov subspace dimension as low as 250 can be enough to Posdziech and Grundmann24 used high-order spectral ele-

TABLE III. Unbounded flow past a cylinder. Convergence of the leading eigenvalue at Re⫽47.08 on M1 with
the Krylov space dimension m and shift parameter ␭.

m ␭⫽0.00⫾0.00i ␭⫽0.25⫾0.00i ␭⫽0.50⫾0.00i

250 ⫺2.499 530⫻10⫺3 ⫾0.716 816 ⫺2.362 255⫻10⫺6 ⫾0.730 913 ⫺2.666 849⫻10⫺6 ⫾0.730 912
500 ⫹3.029 965⫻10⫺3 ⫾0.726 480 ⫺2.668 530⫻10⫺6 ⫾0.730 912 ⫺2.668 473⫻10⫺6 ⫾0.730 912
1000 ⫹2.230 634⫻10⫺4 ⫾0.729 683 ⫺2.668 474⫻10⫺6 ⫾0.730 912 ⫺2.668 473⫻10⫺6 ⫾0.730 912

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
1310 Phys. Fluids, Vol. 16, No. 5, May 2004 M. Sahin and R. G. Owens

FIG. 2. Reciprocal Ritz values for unbounded flow around a circular cylin- FIG. 4. Comparison of drag coefficient versus Reynolds number for un-
der at Re⫽46.76 computed on mesh M3 with Krylov space dimension m bounded flow around a circular cylinder with other results in the literature:
⫽250 and shift parameter ␭⫽0.50⫾0.00i ( ␤ ⫽0.01). 共—兲, numerical results of Henderson 共Ref. 25兲; 共䊊兲, numerical results of
Posdziech and Grundmann 共Ref. 24兲; 共䊐兲, present ( ␤ ⫽0.01, mesh M2兲.

ments the differences between their two sets of results are


due to the use of different blockage ratios in their calcula-
of the lateral boundaries becomes more severe 共see
tions. However, as the Reynolds number increases the differ-
Fornberg,26 for example兲. In addition, Zisis and Mitsoulis27
ence in their computed results becomes smaller. An interest-
showed that the convergence of the total drag at Re⫽0.00
ing convergence study on the extension of the computational
may be very poor as ␤ goes to zero.
domain boundary is given by Posdziech and Grundmann24 at
Re⫽200.00. The authors concluded that the lateral bound-
aries should be set at a distance of at least 70 diameters away IV. FLOW PAST A CONFINED CIRCULAR CYLINDER
in order to obtain a Strouhal number independent of yet „0.1Ë ␤ Ë0.9…
smaller blockage ratios. At lower Reynolds number the effect
Flow around a confined circular cylinder 共as opposed to
the unbounded case兲 is an attractive benchmark problem in
numerical simulation since it does not suffer from any of the
difficulties associated with far-field boundary conditions
共particularly at very low Reynolds numbers兲 and permits the
use of grid points more efficiently in smaller computational
domains. Somewhat surprising, therefore, is that the only
numerical linear stability analysis of Newtonian flow past a
confined cylinder available in the literature would seem to be
that of Chen et al.4 These authors went no further than iden-
tifying the curve of neutral stability for the supercritical Hopf
bifurcation at blockage ratios up to ␤ ⫽0.7. This is regret-
table, because as we shall see in the paragraphs to follow, the
linear stability properties of the flow become rich and there-
fore interesting at higher blockage ratios and Reynolds num-
bers than those considered by Chen et al. In the present study
we consider two-dimensional flow at Reynolds numbers up
to 280 and for blockage ratios in the range 0.1–0.9.

A. Linear stability analysis


The curves of neutral stability computed from the GEVP
FIG. 3. Comparison of Strouhal number versus Reynolds number for un- with a Krylov subspace dimension m⫽250 on mesh M 2 for
bounded flow around a circular cylinder with other results in the literature:
共—兲, experimental work of Williamson 共Ref. 23兲; 共¯兲, numerical results of
␤ 苸 关 0.1,0.9兴 and Re⬍280 are presented in Fig. 5. Our dis-
Henderson 共Ref. 25兲; 共䊊兲, numerical results of Posdziech and Grundmann cussion of these curves will focus on the five distinct curve
共Ref. 24兲; 共䊐兲, present ( ␤ ⫽0.01, mesh M2兲. sections labeled AB, BC, CD, CE, and FG in the same

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
Phys. Fluids, Vol. 16, No. 5, May 2004 A numerical investigation of wall effects 1311

FIG. 5. Change of critical Reynolds number corresponding to both Hopf and pitchfork bifurcations with blockage ratio ␤, computed on M 2. AC: Curve of
neutral stability for Hopf bifurcations about symmetric solution; CD: Transition curve from asymmetric vortex shedding 共smaller ␤兲 to a steady asymmetric
solution 共larger ␤兲; CE: Neutral stability curve for pitchfork bifurcation of steady symmetric solution 共smaller ␤兲 to a steady asymmetric state 共larger ␤兲; FG:
Hopf bifurcation of an asymmetric solution 共smaller ␤兲 to asymmetric vortex shedding 共larger ␤兲. C is a co-dimension 2 point where Hopf and pitchfork
bifurcations occur simultaneously.

TABLE IV. Convergence of critical Reynolds number for different blockage ratios with ␭⫽0.00⫾0.00i.

M1 M2 M3 Chen et al.
Curve section
共see Fig. 5兲 ␤ m Re crit St crit Re crit St crit Re crit St crit Re crit St crit

AC 0.10 500 51.00 0.1206 50.81 0.1210 50.75 0.1211 51.77 0.1116
0.20 250 69.86 0.1559 69.43 0.1566 69.34 0.1567 69.93 0.1559
0.30 250 95.24 0.2079 94.56 0.2090 94.40 0.2093 94.85 0.2085
0.50 250 125.23 0.3369 124.09 0.3393 123.75 0.3399 124.58 0.3382
0.70 250 111.32 0.4714 110.29 0.4752 110.04 0.4762 111.04 0.4744
0.80 250 111.45 0.5324 110.24 0.5363 109.98 0.5374
0.84 250 114.44 0.5530 113.69 0.5568
0.84 250 130.92 0.5510 126.64 0.5557
0.80 250 148.24 0.5324 144.19 0.5383 143.29 0.5398
0.76 250 169.75 0.5115 165.49 0.5186
0.72 250 198.94 0.4872 193.25 0.4955
0.70 250 218.03 0.4737 211.01 0.4827 209.40 0.4851
CE 0.70 250 221.87 216.75 215.53
0.72 250 210.17 205.95
0.76 250 190.65 187.01
0.80 250 173.97 169.49 168.29
0.84 250 161.57 158.15
0.88 250 152.93 149.84
0.90 250 147.78 145.27 144.70
CD 0.68 250 237.33 0.4596 231.06 0.4695
0.66 250 259.55 0.4477 253.08 0.4566
0.64 250 284.56 0.4351 278.01 0.4441
0.62 250 312.66 0.4235 306.27 0.4326
FG 0.82 250 319.80 0.4664
0.82 250 227.44 0.4719
0.84 250 331.02 0.4954
0.84 250 214.00 0.4794 194.30 0.4979
0.88 250 180.43 0.5097 171.28 0.5234
0.90 250 169.44 0.5146 162.82 0.5202 160.50 0.5212

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
1312 Phys. Fluids, Vol. 16, No. 5, May 2004 M. Sahin and R. G. Owens

FIG. 6. Change of base flow with critical Reynolds number and blockage ratio ␤, computed on M 2.

figure. The critical Reynolds numbers and corresponding Up to a blockage ratio ␤ ⫽0.5 Table IV and the neutral
Strouhal numbers 共where appropriate兲 for points on each stability curve AB of Fig. 5 indicate that the flow becomes
curve section and computed on meshes M 1 – M 3 are sup- more stable to two-dimensional infinitesimal disturbances as
plied in Table IV. the blockage ratio increases. All along AB the flow loses
stability to a Hopf bifurcation and the Strouhal number over
1. Curve section AB
this range of blockage ratios is increasing. Between ␤
⫽0.75 and 0.85 it may be seen from Table IV and section
Validation of our numerical stability analysis and direct AB of Fig. 5 that the flow restabilizes slightly leading up to
numerical simulations for flow past an unbounded cylinder point B ( ␤ ⫽ ␤ B ⬇0.855).
␤ ⬇0 has been described in Sec. III. For the confined cylin- In Fig. 6 we show the streamlines of the steady base flow
der problem we have been able to compare our critical Rey- at seven points on the neutral stability curves. Those corre-
nolds and Strouhal numbers for the bifurcation for the sym- sponding to point 1 are typical of those at points on and
metric state with the values for these quantities computed by below curve AB in Fig. 5 where the steady solution is sym-
Chen et al.4 The available results (0.1⭐ ␤ ⭐0.7) of the criti- metric and the only recirculatory region observed is the vor-
cal Reynolds number calculations of Chen et al. are plotted tex pair immediately in the wake of the cylinder itself. That
in Fig. 5 and agreement between our results and theirs over is, for solutions corresponding to parameter space on and
this limited section of the curve AB is excellent. Similarly below AB in Fig. 5 no flow separation on the walls is ob-
excellent agreement in the computed Strouhal numbers was served.
seen over the same range of blockage ratios, both our results
and those of Chen et al. revealing a monotonic increase in 2. Curve sections BC and CE
the critical Strouhal number with the blockage ratio 共see In Fig. 5 section BC represents the part of the critical
Table IV兲. ␤ ⫺Re curve on which the time-dependent state 共symmetric

FIG. 7. Streamlines of unstable symmetric and stable asymmetric solutions at Re⫽150.00 for ␤ ⫽0.9 computed on M3.

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
Phys. Fluids, Vol. 16, No. 5, May 2004 A numerical investigation of wall effects 1313

FIG. 8. Streamlines for the disturbance velocity corresponding to 共a兲 the first and 共b兲 the second leading eigenvectors at Re⫽144.70 for ␤ ⫽0.9 computed
on M3.

periodic oscillations兲, passed into by crossing AB in the di- direction while the eigenvalue on the real axis having largest
rection of increasing Reynolds number, restabilizes to a sym- real part moved right towards the origin. In Fig. 6 we plot the
metric steady state once more. Further increases in the Rey- streamlines on BC at points 2 and 3 to demonstrate how the
nolds number for blockage ratios in the range ␤ C to ␤ B or size of these recirculatory regions as well as their attachment
for choices of ␤ greater than ␤ B may result in the steady distance downstream of the cylinder increase as C is ap-
symmetric solution becoming unstable to two-dimensional proached along the curve BC. In the context of a circular
perturbations via a pitchfork bifurcation into one of two cylinder near a plane boundary such downstream separation
asymmetric states. The curve of neutral stability for this tran- bubbles have been observed both experimentally8 and
sition is labeled CE in Fig. 5. numerically9 for cylinders sufficiently close to the boundary.
The point C is a co-dimension 2 point where Hopf and To gain further insight into the flow transition from
pitchfork bifurcations occur simultaneously. We are able to steady symmetric flow 共between BC and CE and for ␤
estimate the coordinates ( ␤ C ,Re C ) of this point by consid- ⭓ ␤ B ) via a pitchfork bifurcation to steady asymmetric flow
ering it to be the point of intersection of two straight lines 共between DE and FG) we plot in Fig. 7 the streamlines of
drawn through the pairs of points on AC, CD, and EC that
two solutions at a Reynolds number of 150 and blockage
correspond to ␤ ⫽0.68 and 0.7. Since ␤ ⫽0.68 is outside the
ratio of 0.9. It may be seen from Fig. 5 that this point lies
range of blockage ratios corresponding to EC the ordinate
between E (Re⫽144.7,␤ ⫽ ␤ E ⫽0.9) and F (Re⫽160.5,␤
for this value of ␤ is computed to be that at which the lead-
⫽ ␤ F ⫽0.9). Thus, the symmetric solution in the upper plot
ing real eigenvalue in the spectrum of linear perturbations
in Fig. 7 is linearly unstable and the lower plot represents the
about the 共linearly unstable兲 steady symmetric solution is at
the origin. The critical Reynolds numbers at ␤ ⫽0.68 and 0.7 streamlines of one of the stable asymmetric solutions. The
computed on curves AC, CD, and EC are detailed in Table disturbance velocity v in equation 共II兲 is, of course, solenoi-
IV and lead to the estimate ( ␤ C ,Re C )⫽(0.687,224.142). dal. In Fig. 8共a兲 we plot the streamlines associated with the
The occurrence of the transition from a symmetric disturbance field and corresponding to the dominant eigen-
steady state to an asymmetric one on CE is preceded 共in value at E. The addition of a multiple of the eigenvector
Reynolds number兲 by the appearance in the streamlines of a shown in Fig. 8共a兲 to the symmetric steady base flow leads to
pair of downstream separation bubbles on the walls. For ex- one or other of the two asymmetric steady flows, the choice
ample, in the eigenspectrum we observed that for ␤ ⫽0.9 and dependent on the direction of circulation around the symmet-
at Reynolds numbers increasing up to approximately 110 the ric streamlines in Fig. 8共a兲. An anti-clockwise direction leads
complex conjugate pair of leading eigenvalues moved in the to reinforcement of the lower recirculation region and reduc-
direction of the positive real part of the spectrum. At a Rey- tion in the size of the upper bubble. A clockwise direction
nolds number of around 110 separation bubbles appeared on has the opposite effect. The drag coefficient associated with
the walls and with the appearance of the separation bubbles the steady asymmetric solution shown in Fig. 7 is slightly
the leading eigenpair now started to move in the opposite larger than that of the corresponding unstable symmetric one.

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
1314 Phys. Fluids, Vol. 16, No. 5, May 2004 M. Sahin and R. G. Owens

TABLE V. Comparison with the results of Zisis and Mitsoulis 共Ref. 27兲 and Liu et al. 共Ref. 35兲 of computed total drag at Re⫽0.0 for different blockage
ratios ␤.

␤ M1 M2 M3 Extrapolated Zisis and Mitsoulis Liu et al.

0.1 8.9125 8.9100 8.9089 8.9080 8.912 8.9067


0.2 1.6221⫻101 1.6215⫻101 1.6211⫻101 1.6200⫻101 ¯ ¯
0.3 2.7923⫻101 2.7910⫻101 2.7902⫻101 2.7886⫻101 ¯ ¯
0.5 8.8354⫻101 8.8294⫻101 8.8263⫻101 8.8227⫻101 8.8207⫻101 8.8227⫻101
0.7 4.0347⫻102 4.0318⫻102 4.0299⫻102 4.0257⫻102 ¯ ¯
0.9 7.7057⫻103 7.6988⫻103 7.6959⫻103 7.6941⫻103 ¯ ¯

3. Curve section CD ratio decreases the critical Reynolds number.28 In the three-
The section of the neutral stability curves labeled CD in dimensional case, Schrek and Schäfer34 found that fixing the
Fig. 5 represents a transition curve with increasing ␤ from expansion ratio at 1:3 and decreasing the width of the chan-
periodic vortex shedding to the left of this curve 共smaller ␤兲 nel relative to the downstream channel height from ⬁ 共two-
to a steady asymmetric state 共larger ␤兲. At point C the steady dimensional flow兲 through 5 to 2 resulted in a stabilization of
solution is symmetric but moving along the curve CD to- the flow.
wards D causes the growth of one of the recirculatory wall It will be noted from Table IV that for the choices of
regions relative to the other. ␤ ⫽0.82 and 0.84 critical Reynolds numbers of 319.8 and
Since all curves of neutral stability in Fig. 5 have been 331.02, respectively, are added to those that are shown in
determined using a linear stability analysis about a steady Fig. 5. This is to indicate how the curve FG would continue
flow as described in Sec. II we have been unable to plot the if the range of Reynolds numbers were to be extended in Fig.
precise boundaries of the transition region that must exist 5, although at these higher Reynolds numbers it is highly
from symmetric oscillations to asymmetric oscillations as the unlikely that the flow would in reality remain two-
curve CD is approached in parameter space from the left dimensional.
共smaller ␤兲.
B. Direct numerical simulations
4. Curve section FG A few verifications were performed on the results of our
Finally, the steady asymmetric solution of the region be- direct numerical simulations in order to establish their reli-
tween curves DE and FG can become unstable via a Hopf ability. First, Strouhal numbers near the critical Reynolds
bifurcation to asymmetric vortex shedding 共see discussion in numbers corresponding to the onset of periodic vortex shed-
Sec. IV B兲. The transition curve is plotted as FG and the ding and computed with direct numerical simulation were
streamlines of two steady base flows at points 6 and 7 of this found to be in good agreement in a couple of cases with
curve are shown in Fig. 6. In Fig. 8共b兲 we plot the stream- those predicted on the basis of the eigenvalue analysis of
lines of the disturbance velocity corresponding to the leading
complex eigenvalue pair in the spectrum at E. It is the addi-
tion of a mode similar in form to this 共but at a higher Rey-
nolds number兲 that leads to vortex shedding about the asym-
metric state. If the Reynolds number is further increased on
the curve FG additional separation bubbles appear on the
wall further downstream. We also remark that on the curve
FG the separation bubble just behind the cylinder is gener-
ally shorter and more rounded than that computed on the
curves BC and CD.
A strong parallel is thus seen in the present results with
those of numerous other authors 共see, for example, those of
Battaglia et al.,28 Drikakis,29 Fearn et al.,30 Hawa and
Rusak,31 Mishra and Jayaraman,32 and Oliveira33兲 for flows
through both two-dimensional and three-dimensional sym-
metric expansions. All the cited authors report that steady
flow with symmetric recirculatory regions through an expan-
sion geometry encounters a supercritical pitchfork bifurca-
tion at a certain Reynolds number 共dependent, of course, on
the channel geometry兲 and becomes asymmetric. The differ-
ence in the streamwise attachment length of the two recircu-
latory regions 共still in the steady regime兲 becomes larger as
the Reynolds number is further increased from the critical FIG. 9. Computed drag coefficient versus Reynolds number at blockage
value. In the two-dimensional case, increasing the expansion ratios ␤ ⫽0.1, 0.3, 0.5, 0.7, and 0.9.

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
Phys. Fluids, Vol. 16, No. 5, May 2004 A numerical investigation of wall effects 1315

FIG. 10. Change of time-dependent flow with Reynolds number and blockage ratio ␤, computed on M 2.

Sec. IV A. For example, at a blockage ratio of ␤ ⫽0.3 and at Liu et al.35 and the agreement is convincing. It should be
Re⫽100 the corresponding Strouhal number was computed further added that the drag result of Liu et al. for ␤ ⫽0.1 is
from the lift coefficient data over extended time intervals and within 0.007% of the theoretically predicted value of
found to be equal to 0.2115. This compares well with the Faxén.36 The drag coefficient versus Reynolds number is
value of 0.2090 supplied in Table IV and computed at given in Fig. 9 for several blockage ratios. At all the block-
Re crit⫽94.56. Second, we present in Table V results of com- age ratios considered here the drag coefficient behaves like
putations using the three meshes M 1 to M 3 of the drag on 1/Re at low Reynolds numbers. As the blockage ratio in-
the cylinder for various blockage ratios. These are compared creases the range of values of Re over which this remains
with the recent numerical data of Zisis and Mitsoulis27 and true gets smaller.

FIG. 11. Vorticity contours of the periodic flow at Re⫽200.00 and ␤ ⫽0.5 共point 1 in Fig. 10兲 computed on M2. t⫽0 corresponds to the solution having
minimum lift coefficient and the period T⬇2.85.

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
1316 Phys. Fluids, Vol. 16, No. 5, May 2004 M. Sahin and R. G. Owens

FIG. 12. Phase space plots of lift and drag coefficients parametrized with nondimensional time (tU max /D), computed on M2. 共a兲 Re⫽200.00 and ␤ ⫽0.5
共point 1 in Fig. 10兲, 共b兲 Re⫽200.00 and ␤ ⫽0.7 共point 2 in Fig. 10兲, 共c兲 Re⫽200.00 and ␤ ⫽0.9 共point 3 in Fig. 10兲, 共d兲 Re⫽200.00 and ␤ ⫽0.8 共point 4 in
Fig. 10兲, 共e兲 Re⫽160.00 and ␤ ⫽0.8 共point 5 in Fig. 10兲.

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
Phys. Fluids, Vol. 16, No. 5, May 2004 A numerical investigation of wall effects 1317

FIG. 13. Vorticity contours of the periodic flow at Re⫽200.00 and ␤ ⫽0.7 共point 2 in Fig. 10兲 computed on M2. t⫽0 corresponds to the solution having
minimum lift coefficient and the period T⬇2.05.

FIG. 14. Vorticity contours of the periodic solution at Re⫽200.00 and ␤ ⫽0.9 共point 3 in Fig. 10兲 computed on M2. t⫽0 corresponds to the solution having
minimum lift coefficient and the period T⬇1.88.

FIG. 15. Vorticity contours of the periodic solution at Re⫽200.00 and ␤ ⫽0.8 共point 4 in Fig. 10兲 computed on M2. t⫽0 corresponds to the solution having
minimum lift coefficient and the period T⬇1.815.

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
1318 Phys. Fluids, Vol. 16, No. 5, May 2004 M. Sahin and R. G. Owens

FIG. 16. Vorticity contours of the periodic solution at Re⫽160.00 and ␤ ⫽0.8 共point 5 in Fig. 10兲 computed on M2. t⫽0 corresponds to the solution having
minimum lift coefficient and the period T⬇1.806.

In order to elucidate the variation in the critical Rey- been documented by other authors.9 In Fig. 12共a兲 we show
nolds number with blockage ratio observed from the eigen- the C d ⫺C l phase space plot at point 1 of Fig. 10, once fully
value analysis of Sec. IV A we used direct numerical simu- periodic conditions have been established. The average lift is
lation to investigate the wake structure at five different zero and C d and C l are both symmetric in the rising and
locations in the ␤ ⫺Re parameter space and labeled 1–5 in falling parts of each cycle.
Fig. 10. As an aside, and before discussing our results in At a blockage ratio of 0.7 the flow is periodic at a Rey-
detail, we simply note that at blockage ratios ␤ ⬍0.5 direct nolds number of 200.00 共point 2 of Fig. 10兲 with St
numerical simulations revealed that the vortex shedding over ⫽0.4881. The vorticity contours at times t⫽0, T/3 and 2T/3
the cylinder was quite similar to that of the unbounded case, are shown in Figs. 13共a兲–13共c兲, with the period T⬇2.05.
although the vortex street is shorter due to shear in the free Unlike in the case of ␤ ⫽0.5 described in the paragraph
stream. We also note that although blockage effects are ex- above, vortex shedding from the cylinder seems to be almost
pected to delay transition of the cylinder wake to three- suppressed at this Reynolds number, due to the proximity of
dimensional flow, it is possible that at some of the points point 2 to the curve of neutral stability BC. However, there
labeled 1– 4 in Fig. 10 the local velocity is so high that a are very weak vortices shed from both upper and lower lat-
three-dimensional transition occurs for the highly deceler- eral walls. These are well separated from each other and their
ated, separated boundary layers on the channel walls. Verifi- interaction is weak. The phase space plot of the lift and drag
cation of this will have to await fully three-dimensional coefficients at this blockage ratio are shown in Fig. 12共b兲.
simulations, however. Although the time-averaged value of the drag coefficient C d
Time-dependent solutions are presented in Fig. 11 for has increased it is notable that the amplitude of the C d oscil-
␤ ⫽0.5 at a Reynolds number of 200 共point 1 of Fig. 10兲. At lations is an order of magnitude less than that seen at ␤
this Reynolds number the flow has lost its stability to two- ⫽0.5.
dimensional disturbances and has become time-periodic with At a blockage ratio of 0.9 and Re⬎160.5 the flow is
St⫽0.3513, which is higher than for unbounded flow around unsteady and very strong vortices are shed from both the
a circular cylinder (St⫽0.1977). The sequences of three cylinder and the walls. The computed streamlines and vor-
‘‘snapshots’’ in Fig. 11 are taken at times t⫽0, T/3 and 2T/3, ticity contours at Re⫽200.00 共point 3 of Fig. 10兲 are shown
the nondimensional period T being approximately equal to in Fig. 14, each separated from the previous in the series by
2.85 and determined from the lift coefficient data over long a third of a period T/3. At this Reynolds number the flow is
time periods. t⫽0 corresponds to a minimum in the lift co- periodic with St⫽0.5314 (T⬇1.88) and this compares rea-
efficient once fully periodic vortex shedding is established. It sonably well 共see Table IV兲 with the Strouhal number of
may be seen from Figs. 11共a兲–11共c兲 that vortex shedding 0.5202 computed from the GEVP on mesh M 2 at the critical
occurs both from the cylinder and the channel walls. As these Reynolds number Re crit⫽162.82. Vortices having the same
vortices move downstream the trajectories of clockwise vor- sign merge just behind the cylinder and are then transported
tices shed from the upper part of the cylinder cross those downstream. However, the vortex street formed behind the
having opposite sign 共and shed from the lower part of the cylinder is quite different from the well-known von Kármán
cylinder兲 so that wall proximity effects are seen to give rise street in that very strong opposite-sign vortices with smaller
to a reverse von Kármán street. The same phenomenon has structure move downstream and interact with the wall, cre-

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
Phys. Fluids, Vol. 16, No. 5, May 2004 A numerical investigation of wall effects 1319

ating strong vortices there. Although the streamlines of the V. CONCLUSIONS


eigenvector shown in Fig. 8共b兲 will be slightly modified by In this paper we have computed with greater accuracy
the time the Reynolds number reaches 200, the size of the and over a larger range of blockage ratios than has proved
cellular structures in the wake of the eigenvector are close to possible in the past the effects on the drag and linear stability
what is seen in the direct numerical solution for the vorticity of lateral wall proximity for flow past a cylinder at Reynolds
in Fig. 14. Oscillations are now about an asymmetric state numbers up to 280.
and lead to drastic increases in both the lift and drag coeffi- Some of the rich and complex dynamics of the system
cient values. Their variation over a cycle with dimensionless for sufficiently high Reynolds numbers and blockage ratios
time is shown in Fig. 12共c兲. In this figure it may be seen that have been uncovered and discussed. In particular, we have
not only is the time-averaged value of C d greater than for the found that for Re⭐280 and ␤ ⭐0.9 there are 共at least兲 three
two previous blockage ratios considered, but the amplitudes separate curves of neutral stability: 共a兲 Hopf bifurcation of a
of oscillation of both coefficients has dramatically increased symmetric state, 共b兲 pitchfork bifurcation of a symmetric
after a tendency observed up to ␤ ⬇0.75 of successively di- state to one of two asymmetric states, and 共c兲 Hopf bifurca-
minishing amplitudes. We also note for the first time that the tion of an asymmetric state leading to asymmetric oscilla-
lift coefficient is no longer symmetric in the rising and fall- tions thereafter. In addition, we have drawn attention to a
transition region from symmetric vortex shedding to asym-
ing parts of each cycle 共the ‘‘figure of 8’’ is distorted兲, al-
metric vortex shedding with increasing blockage ratio. Fur-
though the average value of C l over one cycle is zero.
ther increases in the blockage ratio 共crossing CD in Fig. 10兲
The flow behavior in the parametric region between
leads to restabilization to a steady asymmetric solution.
curves BD and FG is particularly interesting since although A co-dimension 2 point where pitchfork and Hopf bifur-
the steady base flow solutions are linearly stable they appear cations occur simultaneously has been identified and a region
to be finitely unstable. Although our primary concern in this in parameter space 共either side of the locus of the pitchfork
paper is the study of wall effects on the linear stability of bifurcation兲 seems to exist where the steady solution is lin-
flow past a confined cylinder, in Fig. 10 we plot the instan- early stable but unstable to finite two-dimensional perturba-
taneous streamlines of two pairs of different solutions, both tions.
pairs being for a blockage ratio ␤ ⫽0.8 but corresponding to
two different values of Re: at point 4 Re⫽200 whereas at ACKNOWLEDGMENTS
point 5 Re⫽160. Shown in the uppermost plots at points 4
The authors wish to thank Peter Monkewitz for sharing
and 5 are the steady and linearly stable solutions, but these
with them interesting and illuminating insights into the pri-
may be sent into permanently unsteady states by running, for
mary and secondary instability mechanisms. The work of the
example, the time-dependent code with the geometrically
first author was supported by the Swiss National Science
rescaled steady base flow of point 2 ( ␤ ⫽0.7,Re⫽200) as Foundation, Grant No. 21-61865.00.
the initial ‘‘guess.’’ Snapshots of unsteady flow at times t
⫽0, T/3, and 2T/3 at point 4 are presented in Figs. 15共a兲– 1
C. H. K. Williamson, ‘‘Vortex dynamics in the cylinder wake,’’ Annu. Rev.
15共c兲, with the period T⬇1.815. Unlike the other unsteady Fluid Mech. 28, 477 共1996兲.
2
cases considered so far in this paper, the recirculation regions P. Anagnostopoulos, G. Iliadis, and S. Richardson, ‘‘Numerical study of
on the upper wall are much larger than those on the lower the blockage effect on viscous flow past a circular cylinder,’’ Int. J. Numer.
Methods Fluids 22, 1061 共1996兲.
wall and they can move far downstream. Additionally, the 3
M. Coutanceau and R. Bouard, ‘‘Experimental determination of the main
recirculatory region at the lower part of the cylinder is larger features of the viscous flow in the wake of a circular cylinder in uniform
than that at the upper part. This asymmetry may be seen from translation. Part 1. Steady flow,’’ J. Fluid Mech. 79, 231 共1977兲.
4
J.-H. Chen, W. G. Pritchard, and S. J. Tavener, ‘‘Bifurcation for flow past
the phase space plot in Fig. 12共d兲. The time average of the a cylinder between parallel planes,’’ J. Fluid Mech. 284, 23 共1995兲.
5
lift coefficient is no longer zero and the lift curve is asym- M. Behr, S. Hastreiter, S. Mittal, and T. E. Tezduyar, ‘‘Incompressible flow
past a circular cylinder: Dependence of the computed flow field on the
metric in the rising and falling part of each cycle. However,
location of the lateral boundaries,’’ Comput. Methods Appl. Mech. Eng.
if the Reynolds number is chosen equal to 160 共point 5 in 123, 309 共1995兲.
6
Fig. 10兲 the flow becomes symmetric again since point 5 is P. K. Stansby and A. Slaouti, ‘‘Simulation of vortex shedding including
below the curve CE. The computed vorticity contours are blockage by the random-vortex and other methods,’’ Int. J. Numer. Meth-
ods Fluids 17, 1003 共1993兲.
given in Fig. 16 with T⬇1.806. At this point the flow struc- 7
C. Lei, L. Cheng, and K. Kavanagh, ‘‘Re-examination of the effect of a
ture is quite similar to that of point 2 with vortex shedding plane boundary on force and vortex shedding of a circular cylinder,’’ J.
from the upper and lower walls. The lift and drag coefficients Wind. Eng. Ind. Aerodyn. 80, 263 共1999兲.
8
P. W. Bearman and M. M. Zdravkovich, ‘‘Flow around a circular cylinder
which are supplied in the phase space plot in Fig. 12共e兲 are near a plane boundary,’’ J. Fluid Mech. 89, 33 共1978兲.
9
also similar to those of point 2 with zero time average of the L. Zovatto and G. Pedrizzetti, ‘‘Flow around a circular cylinder between
lift coefficient. parallel walls,’’ J. Fluid Mech. 440, 1 共2001兲.
10
C. P. Jackson, ‘‘A finite-element study of the onset of vortex shedding in
Numerical experiments at a blockage ratio of 0.9 and a flow past variously shaped bodies,’’ J. Fluid Mech. 182, 23 共1987兲.
11
Reynolds number of 500 indicated that the flow had become M. Sahin, ‘‘Solution of the incompressible unsteady Navier–Stokes equa-
chaotic. However, it seems unlikely that the flow is still two- tions only in terms of the velocity components,’’ Int. J. Comput. Fluid
Dyn. 17, 199 共2003兲.
dimensional at this Reynolds number and presentation of our 12
M. Sahin and R. G. Owens, ‘‘A novel fully-implicit finite volume method
results will have to await a fully three-dimensional analysis. applied to the lid-driven cavity problem. Part I. High Reynolds number

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions
1320 Phys. Fluids, Vol. 16, No. 5, May 2004 M. Sahin and R. G. Owens

flow calculations,’’ Int. J. Numer. Methods Fluids 42, 57 共2003兲. 24


O. Posdziech and R. Grundmann, ‘‘Numerical simulation of the flow
13
M. Sahin and R. G. Owens, ‘‘A novel fully-implicit finite volume method around an infinitely long circular cylinder in the transition regime,’’ Theor.
applied to the lid-driven cavity problem. Part II. Linear stability analysis,’’ Comput. Fluid Dyn. 15, 121 共2001兲.
Int. J. Numer. Methods Fluids 42, 79 共2003兲. 25
R. D. Henderson, ‘‘Detail of the drag curve near the onset of vortex shed-
14
A. Kourta, M. Braza, P. Chassaing, and H. Haminh, ‘‘Numerical analysis ding,’’ Phys. Fluids 7, 2102 共1995兲.
26
of a natural and excited two-dimensional mixing layer,’’ AIAA J. 25, 279 B. Fornberg, ‘‘A numerical study of steady viscous flow past a circular
共1987兲. cylinder,’’ J. Fluid Mech. 98, 819 共1980兲.
15 27
G. Jin and M. Braza, ‘‘A nonreflecting outlet boundary condition for in- Th. Zisis and E. Mitsoulis, ‘‘Viscoplastic flow around a cylinder kept
compressible unsteady Navier–Stokes calculations,’’ J. Comput. Phys. between parallel plates,’’ J. Non-Newtonian Fluid Mech. 105, 1 共2002兲.
107, 239 共1993兲. 28
F. Battaglia, S. J. Tavener, A. K. Kulkarni, and C. L. Merkle, ‘‘Bifurcation
16
W. E. Arnoldi, ‘‘The principle of minimized iterations in the solution of of low Reynolds number flows in symmetric channels,’’ AIAA J. 35, 99
the matrix eigenvalue problem,’’ Q. Appl. Math. 9, 17 共1951兲. 共1997兲.
17 29
Y. Saad, ‘‘Variations on Arnoldi’s method for computing eigen elements of D. Drikakis, ‘‘Bifurcation phenomena in incompressible sudden expansion
large unsymmetric matrices,’’ Linear Algebr. Appl. 34, 269 共1980兲. flows,’’ Phys. Fluids 9, 76 共1997兲.
18 30
R. Natarajan, ‘‘An Arnoldi-based iterative scheme for nonsymmetric ma- R. M. Fearn, T. Mullin, and K. A. Cliffe, ‘‘Nonlinear flow phenomena in a
trix pencils arising in finite element stability problems,’’ J. Comput. Phys. symmetric sudden expansion,’’ J. Fluid Mech. 211, 595 共1990兲.
100, 128 共1992兲. 31
T. Hawa and Z. Rusak, ‘‘The dynamics of a laminar flow in a symmetric
19
P. R. Amestoy, I. S. Duff, and J.-Y. L’Excellent, ‘‘Multifrontal parallel channel with a sudden expansion,’’ J. Fluid Mech. 436, 283 共2001兲.
32
distributed symmetric and unsymmetric solvers,’’ Comput. Methods Appl. S. Mishra and K. Jayaraman, ‘‘Asymmetric flows in planar symmetric
Mech. Eng. 184, 501 共2000兲. channels with large expansion ratio,’’ Int. J. Numer. Methods Fluids 38,
20
P. R. Amestoy, I. S. Duff, J. Koster, and J.-Y. L’Excellent, ‘‘A fully asyn- 945 共2002兲.
33
chronous multifrontal solver using distributed dynamic scheduling,’’ P. J. Oliveira, ‘‘Asymmetric flows of viscoelastic fluids in symmetric pla-
SIAM J. Matrix Anal. Appl. 23, 15 共2001兲. nar expansion geometries,’’ J. Non-Newtonian Fluid Mech. 114, 33
21
J. L. Steger and R. L. Sorenson, ‘‘Automatic mesh-point clustering near a 共2003兲.
34
boundary in grid generation with elliptic partial differential equations,’’ J. E. Schreck and M. Schäfer, ‘‘Numerical study of bifurcation in three-
Comput. Phys. 33, 405 共1979兲. dimensional sudden channel expansions,’’ Comput. Fluids 29, 583 共2000兲.
22 35
Y. Ding and M. Kawahara, ‘‘Three-dimensional linear stability analysis of A. W. Liu, D. E. Bornside, R. C. Armstrong, and R. A. Brown, ‘‘Viscoelas-
incompressible viscous flows using the finite element method,’’ Int. J. tic flow of polymer solutions around a periodic, linear array of cylinders:
Numer. Methods Fluids 31, 451 共1999兲. comparisons of predictions for microstructure and flow fields,’’ J. Non-
23
C. H. K. Williamson, ‘‘Oblique and parallel modes of vortex shedding in Newtonian Fluid Mech. 77, 153 共1998兲.
36
the wake of a circular cylinder at low Reynolds numbers,’’ J. Fluid Mech. O. H. Faxén, ‘‘Forces exerted on a rigid cylinder in a viscous fluid be-
206, 579 共1989兲. tween two parallel fixed planes,’’ R. Swed. Acad. Eng. Sci. 187, 1 共1946兲.

Downloaded 31 Jul 2013 to 66.194.72.152. This article is copyrighted as indicated in the abstract. Reuse of AIP content is subject to the terms at: http://pof.aip.org/about/rights_and_permissions

You might also like