flow of newtonian and power-law fluids past an elliptical ...hron/nmod041/ie100251w.pdf · flow of...

13
Flow of Newtonian and Power-Law Fluids Past an Elliptical Cylinder: A Numerical Study P. Koteswara Rao, Akhilesh K. Sahu, and R. P. Chhabra* Department of Chemical Engineering, Indian Institute of Technology, Kanpur 208016, India Extensive numerical simulations of the 2-D laminar flow of power-law fluids over elliptical cylinders with different aspect ratios have been carried out to establish the conditions for the onset of wake formation and the onset of vortex shedding. The continuity and momentum equations were solved numerically using FLUENT (version 6.3.26). The influence of the power-law index (0.3 e n e 1.8) and the aspect ratio (E ) b/a; 0.2 e E e 5) of the cylinder on the critical values of the Reynolds number denoting the onsets of flow separation and vortex shedding are presented. For shear-thinning (n < 1) fluid behavior, the onsets of wake formation and vortex shedding are both seen to be postponed to higher Reynolds numbers as compared to those in shear-thickening fluids (n > 1). Also, the values of the Strouhal number (St c ) and the time-average drag coefficient (C j D ) corresponding to the cessation of the steady-flow regime are presented. Velocity vector plots denoting the flow separation and vorticity profiles showing the vortex shedding are also included. The delineation of different flow regimes also helps identify the range of validity of some of the results on flow and heat transfer available in the literature. 1. Introduction The flow of fluids past cylinders of various cross sections represents an idealization of several industrially important applications. Typical examples include the flow in tubular and pin-type heat exchangers; filtration screens used for clarifying suspensions, sludges, and polymer melts; membrane-based separation modules; support structures exposed to flow streams of liquids; formation of weld lines in polymer processing; and thermal processing of variously shaped food particles. In addition to the foregoing pragmatic significance, there is an intrinsic theoretical interest in such model flows to further our understanding of the underlying physical processes. Therefore, bluff body flows constitute an important class of problems within the domain of fluid mechanics. Consequently, over the past 100 years or so, a voluminous body of information encompassing wide-ranging phenomena associated with flow past a cylinder has accrued; see, for example, refs 1-3. A quick inspection of these reviews clearly shows that the bulk of the literature relates to the flow of Newtonian fluids past a circular cylinder. Cylinders of noncircular cross sections, notably elliptical, are now receiving increasing attention because of the enhanced performance of tubular heat exchangers (low pressure drop and/ or high heat transfer rate) fitted with elliptical tubes as opposed to circular tubes. Additional examples are found in aerosol filters 4-6 and the use of elliptical cylinders as a model for artificial lungs. 7 Admittedly, the flow in most of the aforemen- tioned settings involving multiple cylinders is very complex; it is readily acknowledged that a systematic study of the flow past a single cylinder not only provides valuable insights into the nature of flow, but also serves as a useful starting point to understand the flow in real-life multicylinder applications. Similarly, many materials encountered in industrial practice, notably multiphase mixtures (foams, dispersions, emulsions), polymeric melts and solutions, and soap solutions, among others, display a range of rheological complexities such as shear thinning, shear thickening, yield stress, and viscoelasticity. 8 Despite their wide occurrence in diverse engineering applica- tions, very little information is available on the flow of power- law-type non-Newtonian fluids past an elliptical cylinder. This work is thus concerned with the flow of Newtonian and power- law fluids over an elliptical cylinder of various aspect ratios. It is useful to recall here that, even for the flow of Newtonian fluids over a circular cylinder, the flow transits from one flow regime to another depending on the value of the Reynolds number. Undoubtedly, the changes in the flow pattern directly impinge on the nature of the flow at both microscopic and macroscopic levels. Thus, at very low Reynolds numbers, the flow is laminar and steady and does not separate from the surface of the cylinder. This flow regime persists up to about Re ) 6, and a visible wake appears at Re ) 6.5, although the flow is still steady and laminar. 9 With a gradual increase in the value of the Reynolds number, the wake region grows and ultimately becomes asymmetric and periodic in time at Re 45-50, thereby leading to the onset of vortex shedding. Thus, for an unconfined circular cylinder, the transitions between different flow regimes are adequately described in terms of the value of the Reynolds number. The corresponding values of the critical Reynolds numbers for the flow of power-law fluids have been reported only recently. 9 Depending on the values of the flow behavior index, the flow can remain attached to the surface of the cylinder; for example, it does not separate up to about Re 11-12 in a highly shear-thinning fluid, whereas the flow transits to the periodic-flow regime at Re 33-34 in a highly shear-thickening fluid. It is somewhat surprising that very little analogous information is available for elliptical cylinders, even in Newtonian fluids, let alone in power-law fluids. This work thus aims to fill this gap in the current literature. The delineation of the different flow regimes directly influences the choice of numerics, such as whether to use a steady or an unsteady solver. In particular, reported herein are the values of the critical Reynolds numbers denoting the onset of the separation of flow and the onset of vortex shedding as functions of the power-law index and aspect ratio of the cylinder. However, prior to undertaking a detailed presentation of the new results, it is instructive to summarize the salient features of the scant literature available on the flow of Newtonian and * To whom correspondence should be addressed. Tel.: +91-512- 259 7393. Fax: +91-512-259 0104. E-mail: [email protected]. Ind. Eng. Chem. Res. 2010, 49, 6649–6661 6649 10.1021/ie100251w 2010 American Chemical Society Published on Web 06/21/2010

Upload: vomien

Post on 27-Mar-2018

218 views

Category:

Documents


2 download

TRANSCRIPT

Flow of Newtonian and Power-Law Fluids Past an Elliptical Cylinder:A Numerical Study

P. Koteswara Rao, Akhilesh K. Sahu, and R. P. Chhabra*

Department of Chemical Engineering, Indian Institute of Technology, Kanpur 208016, India

Extensive numerical simulations of the 2-D laminar flow of power-law fluids over elliptical cylinders withdifferent aspect ratios have been carried out to establish the conditions for the onset of wake formation andthe onset of vortex shedding. The continuity and momentum equations were solved numerically using FLUENT(version 6.3.26). The influence of the power-law index (0.3 e n e 1.8) and the aspect ratio (E ) b/a; 0.2 eE e 5) of the cylinder on the critical values of the Reynolds number denoting the onsets of flow separationand vortex shedding are presented. For shear-thinning (n < 1) fluid behavior, the onsets of wake formationand vortex shedding are both seen to be postponed to higher Reynolds numbers as compared to those inshear-thickening fluids (n > 1). Also, the values of the Strouhal number (Stc) and the time-average dragcoefficient (CjD) corresponding to the cessation of the steady-flow regime are presented. Velocity vector plotsdenoting the flow separation and vorticity profiles showing the vortex shedding are also included. Thedelineation of different flow regimes also helps identify the range of validity of some of the results on flowand heat transfer available in the literature.

1. Introduction

The flow of fluids past cylinders of various cross sectionsrepresents an idealization of several industrially importantapplications. Typical examples include the flow in tubular andpin-type heat exchangers; filtration screens used for clarifyingsuspensions, sludges, and polymer melts; membrane-basedseparation modules; support structures exposed to flow streamsof liquids; formation of weld lines in polymer processing; andthermal processing of variously shaped food particles. Inaddition to the foregoing pragmatic significance, there is anintrinsic theoretical interest in such model flows to further ourunderstanding of the underlying physical processes. Therefore,bluff body flows constitute an important class of problems withinthe domain of fluid mechanics. Consequently, over the past 100years or so, a voluminous body of information encompassingwide-ranging phenomena associated with flow past a cylinderhas accrued; see, for example, refs 1-3. A quick inspection ofthese reviews clearly shows that the bulk of the literature relatesto the flow of Newtonian fluids past a circular cylinder.Cylinders of noncircular cross sections, notably elliptical, arenow receiving increasing attention because of the enhancedperformance of tubular heat exchangers (low pressure drop and/or high heat transfer rate) fitted with elliptical tubes as opposedto circular tubes. Additional examples are found in aerosolfilters4-6 and the use of elliptical cylinders as a model forartificial lungs.7 Admittedly, the flow in most of the aforemen-tioned settings involving multiple cylinders is very complex; itis readily acknowledged that a systematic study of the flow pasta single cylinder not only provides valuable insights into thenature of flow, but also serves as a useful starting point tounderstand the flow in real-life multicylinder applications.Similarly, many materials encountered in industrial practice,notably multiphase mixtures (foams, dispersions, emulsions),polymeric melts and solutions, and soap solutions, among others,display a range of rheological complexities such as shearthinning, shear thickening, yield stress, and viscoelasticity.8

Despite their wide occurrence in diverse engineering applica-

tions, very little information is available on the flow of power-law-type non-Newtonian fluids past an elliptical cylinder. Thiswork is thus concerned with the flow of Newtonian and power-law fluids over an elliptical cylinder of various aspect ratios.

It is useful to recall here that, even for the flow of Newtonianfluids over a circular cylinder, the flow transits from one flowregime to another depending on the value of the Reynoldsnumber. Undoubtedly, the changes in the flow pattern directlyimpinge on the nature of the flow at both microscopic andmacroscopic levels. Thus, at very low Reynolds numbers, theflow is laminar and steady and does not separate from the surfaceof the cylinder. This flow regime persists up to about Re ) 6,and a visible wake appears at Re ) 6.5, although the flow isstill steady and laminar.9 With a gradual increase in the valueof the Reynolds number, the wake region grows and ultimatelybecomes asymmetric and periodic in time at Re ≈ 45-50,thereby leading to the onset of vortex shedding. Thus, for anunconfined circular cylinder, the transitions between differentflow regimes are adequately described in terms of the value ofthe Reynolds number. The corresponding values of the criticalReynolds numbers for the flow of power-law fluids have beenreported only recently.9 Depending on the values of the flowbehavior index, the flow can remain attached to the surface ofthe cylinder; for example, it does not separate up to about Re≈ 11-12 in a highly shear-thinning fluid, whereas the flowtransits to the periodic-flow regime at Re ≈ 33-34 in a highlyshear-thickening fluid. It is somewhat surprising that very littleanalogous information is available for elliptical cylinders, evenin Newtonian fluids, let alone in power-law fluids.

This work thus aims to fill this gap in the current literature.The delineation of the different flow regimes directly influencesthe choice of numerics, such as whether to use a steady or anunsteady solver. In particular, reported herein are the values ofthe critical Reynolds numbers denoting the onset of theseparation of flow and the onset of vortex shedding as functionsof the power-law index and aspect ratio of the cylinder.However, prior to undertaking a detailed presentation of thenew results, it is instructive to summarize the salient featuresof the scant literature available on the flow of Newtonian and

* To whom correspondence should be addressed. Tel.: +91-512-259 7393. Fax: +91-512-259 0104. E-mail: [email protected].

Ind. Eng. Chem. Res. 2010, 49, 6649–6661 6649

10.1021/ie100251w 2010 American Chemical SocietyPublished on Web 06/21/2010

power-law fluids over elliptical cylinders, which, in turn,facilitates the presentation of new results obtained in this work.

2. Previous Work

The flow of Newtonian fluids over elliptical cylinders hasreceived much less attention than that accorded to the case ofa circular cylinder. Early studies10-13 were based on theapplication of matched asymptotics to obtain analytical solutionof Oseen’s linearized equation that led to the prediction ofvelocity and pressure fields close to the elliptical cylinder.Although it is impossible to delineate the range of validity ofthese solutions in a rigorous fashion, these analyses are believedto be useful up to about Re ≈ 1-2. The next generation ofsolutions was based on numerical solutions of the Navier-Stokesequations. For instance, Epstein and Masliyah14 presented thesteady-state results up to Re ) 90 for elliptical cylinders withaspect ratios (E ) b/a) ranging from 0.2 to 1. Meller15

numerically solved the elliptical transformation of the vorticity-stream function formulation of the Navier-Stokes equationsusing the finite difference method and matrix pivotal condensa-tion method for two values of the Reynolds numbers (i.e., Re) 20 and 40) and for an elliptical cylinder with an aspect ratioof 10 at an angle of incidence of 30°. D’Alessio and Dennis16

presented drag and lift coefficients for Re ) 5 and 20 fordifferent inclinations of the elliptical cylinder ranging from 0°to 90°. Subsequently, Dennis and Young17 studied the impactof the angle of incidence (0-90°) on the flow characteristics,but their results were limited to the so-called steady-flow regime,1 e Re e 40, and/or for a single value of the aspect ratio ofcylinder, namely, 5. Notwithstanding the differences arising fromthe different numerics employed, broadly speaking, the valuesof the drag coefficient and pressure distribution on the surfaceof the cylinder are consistent with each other for Newtonianfluids.

The unsteady flow of an abruptly started elliptical cylinderat 45° incidence was studied by Lugt and Haussling.18 The majorthrust of their study was the prediction of the time required forthe flow to reach steady or quasisteady state by following theevolution of streamlines and drag and lift coefficients with time.Subsequently, Patel19 revisited this flow configuration for a rangeof angles of incidence (0°, 30°, 45°, 90°) for a fixed value ofRe ) 200. Using the spectral element method, Johnson et al.20

investigated the time-dependent flow over an elliptical cylinderover the range of conditions characterized by 30 e Re e 200and for aspect ratios varying from 0.01 (vertical thin plate) to1 (circular cylinder). They reported the critical values of theReynolds number denoting the cessation of the steady-flowregime to be only weakly dependent on the aspect ratio of thecylinder over this range of aspect ratios. For aspect ratios lessthan 1, the steady-flow regime was observed to persist up toabout Re ≈ 35-45, which is obviously due to the reducedstreamlining of the cylinder, albeit this value is only slightlylower than the oft-quoted value of 46-48 for a circular cylinder.This clearly casts some doubts on the validity of the steady-flow assumption up to Re ) 90 in an earlier study.14 Unsteadyflow over an elliptical cylinder (aspect ratios 0.5 and 0.6) up toRe ≈ 500 was investigated by Ahmad and Badr.21 Theyexamined the unsteadiness arising from two sources, namely,the fluctuating imposed flow and the vortex-shedding phenom-enon. Aside from the aforementioned studies, which utilizedthe complete Navier-Stokes equations, Khan et al.22 employedthe boundary-layer approximation to develop closed-formanalytic expressions for skin friction and heat transfer from anelliptical cylinder. It is fair to say that, with the exception of

the work of Johnson et al.,20 none of the studies mentionedabove addressed the delineation of flow regimes for the flowpast an unconfined elliptical cylinder. However, there have beena few studies that have attempted to elucidate the influence ofthe aspect ratio of the cylinder on the transition from the steady-flow regime to a time-dependent periodic-flow regime. Forinstance, Jackson23 carried out a finite-element-based numericalstudy to delineate the transitional Reynolds number denotingthe onset of the vortex-shedding flow regime. In particular, fora cylinder with an aspect ratio of 0.5, he reported the criticalvalue of the Reynolds number to vary from 35.7 to 141 as theangle of orientation of cylinder to flow changed from 0° to 90°.Similarly, Johnson et al.20 reported on the various types ofshedding patterns observed with elliptical cylinders with aspectratios less than 1. More recently, Faruquee et al.24 investigatedthe first transition denoting the onset of flow separation forelliptical cylinders. They fixed the Reynolds number (based onthe hydraulic diameter of the cylinder) at 40 and found that thefirst signature of flow separation occurred at about the aspectratio of 0.34. Based on their24 numerical results, they alsopresented an empirical relation between the wake length andthe aspect ratio at a fixed value of Reynolds number, and thewake size was seen to increase with increasing aspect ratio (e1).This issue was investigated in a more detailed manner by Stackand Bravo,25 who studied the onset of flow separation for a flatplate (aspect ratio of 0), circular cylinder (aspect ratio of 1)using a cellular automata model and FLUENT. For fixed valuesof the Reynolds number, they varied the value of the aspectratio to locate the onset of flow separation. All in all, it is thusfair to say that only limited information is available on the firsttwo transitions, namely, flow separation/no separation andsteady/time-dependent periodic-flow regimes for elliptical cyl-inders even in Newtonian fluids.

In contrast, the analogous body of knowledge for power-lawfluids is not only very limited but also of recent vintage evenfor circular cylinders. In view of the available information onthe critical Reynolds numbers corresponding to the first twotransitions,9 suffice it to say here that reliable results on dragand wake characteristics are now available on the flow of power-law fluids past a circular cylinder in the steady-flow regime.26-30

The role of planar confining walls has been examined by Bhartiet al.31 Limited results are also available on convective heattransfer from a circular cylinder submerged in power-law fluidsin the steady-flow regime.32-34 Broadly, all else being equal,shear-thinning fluid behavior promotes heat transfer, and shearthickening has deleterious effect on it.32-34 Further enhance-ments in heat-transfer rates are possible in the presence ofbuoyancy effects.33,34 More recently, Patnana et al.35,36 carriedout a numerical study in the periodic-flow regime to elucidatethe role of power-law rheology in the vortex-shedding and heat-transfer characteristics of a cylinder in power-law fluids. Insummary, whereas a reasonable body of knowledge is nowavailable on the momentum and heat-transfer characteristics fora circular cylinder in power-law fluids in the steady-flow regime,only limited information is available in the time-dependentperiodic-flow regime.

Finally, to the best of our knowledge, only two previousstudies have been reported on the flow of power-law fluids overelliptical cylinders.37,38 Sivakumar et al.37 and Bharti et al.38

presented extensive results on flow (drag, wake size, surfacepressure, and vorticity profiles) and heat-transfer (Nusseltnumber) over the range of parameters expressed as 0.01 e Ree 40, 0.2 e E e 5, and 0.2 e n e 1.8, thereby covering bothshear-thinning and shear-thickening fluid behaviors. The role

6650 Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010

of the power-law index becomes accentuated depending on thevalue of the aspect ratio of the cylinder. However, the flow wasassumed to be steady a priori over these ranges of conditions,and this is indeed the main weakness of their work. Therefore,the utility of their results rests on the validity of this assumption,which necessitates the delineation of the critical Reynoldsnumbers marking the limit of the steady-flow regime for variousvalues of aspect ratio and power-law index.

Based on the preceding discussion, it is thus safe to concludethat only limited information is available on the transitionalReynolds numbers for elliptical cylinders even in Newtonianmedia and no prior results are available for power-law fluids.This work reports on the critical values of the Reynolds numberdenoting the onsets of flow separation and vortex shedding foran elliptical cylinder over a wide range of power-law indices(0.3 e n e 1.8) and cylinder aspect ratios (0.2 e E e 5). Thisinformation is useful not only in its own right, but also toascertain the range of applicability of the currently availableresults in the literature.

3. Problem Statement and Governing Equations

Let us consider the two-dimensional (2-D) unconfined flowof an incompressible power-law fluid (uniform approach velocityU∞) over a long elliptical cylinder of aspect ratio E orientednormal to the oncoming uniform flow. Here, the aspect ratio of

the cylinder is defined as the ratio of the length of the axisparallel to the flow to the length of the axis perpendicular tothe flow. Thus, for cases with E > 1, it becomes the major axisdivided by the minor axis, whereas for cases with E < 1, it isthe minor axis divided by the major axis. The unconfined flowcondition is approximated here by enclosing the ellipticalcylinder in a large circular outer boundary of diameter D∞, asshown schematically in Figure 1. A prudent choice of D∞ is bynecessity a tradeoff between the computational effort (whichincreases rapidly with increasing value of D∞) and the extentof boundary effects on the accuracy of the results (whichdeteriorates with decreasing value of D∞).

The continuity and momentum equations for this flow arewritten as

where F,U, f, and σ are the density, velocity vector, body force,and stress tensor, respectively. For incompressible fluids, thestress tensor is made up of two components, the isotropicpressure and the extra stress tensor τ, i.e.

Figure 1. Schematic representation of unconfined flow past an elliptical cylinder (E ) b/a).

Continuity equation∇ ·U ) 0 (1)

Momentum equation

F(DUDt

- f) - ∇ ·σ ) 0 (2)

Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010 6651

The extra stress tensor, τ, is, in turn, related to the rate ofdeformation tensor ε(u) through the scalar viscosity of the fluidas

where ε(u), the rate of deformation tensor, is related to thevelocity field as

and the power-law viscosity is given by

where I2 is the second invariant of the rate of deformation tensorand n is the flow behavior index. Evidently, n < 1 denotes shear-thinning behavior, n > 1 corresponds to shear-thickeningbehavior, and n ) 1 represents the standard Newtonian fluidbehavior. The expressions for I2 in terms of velocity and velocitygradients are available in standard text books; for example, seeBird et al.39

To complete the problem statement, appropriate boundaryconditions at the inlet and outlet boundaries, which are locatedat the front half and rear half, respectively, of the computationaldomain (Figure 1A), and at the cylinder wall must be specified.For this flow, these conditions are written as follows:

(1) At the inlet boundary, the condition of uniform velocityin the x direction is prescribed by

(2) On the surface of the cylinder, because it is a solid surface,the standard no-slip condition is used, that is

(3) At the exit boundary, the default outflow boundarycondition in FLUENT is used, which assumes a zero diffusionflux for all flow variables (i.e., ∂φ/∂x ) 0, where φ is a scalarvariable such as Ux or Uy), and this is similar to the fullydeveloped flow condition. It is important to note here thatgradients in the cross-stream direction might still exist at theoutflow boundary. Because the flow becomes steady at the exitboundary as a result of the dissipation of vortices at the outerdomain, this condition can be prescribed for both the steadyand vortex-shedding flow regimes near the cylinder as

(4) At the plane of symmetry (y ) 0), prior to the onset ofthe vortex-shedding regime, the flow is symmetric about themidplane. Hence, it is sufficient to use half of the flow domainfor steady simulations for estimating the critical Reynoldsnumber for the onset of flow separation, namely

The numerical solution of eqs 1 and 2 subject to the boundaryconditions in eqs 7-10 maps the flow domain in terms of thevelocity and pressure fields, and these, in turn, are used to

compute global characteristics such as the drag and lift coef-ficients and Strouhal number.

At low Reynolds numbers, the flow is expected to be steadyand symmetric, and the scaling considerations suggest that theflow is governed by four dimensionless parameters, namely,aspect ratio, power-law index, Reynolds number, and dragcoefficient, with both lift coefficients and Strouhal number beingzero under these conditions. In the time-dependent periodic-flow regime, the Strouhal number and lift coefficient achievenonzero values and are likely to be functions of the aspect ratio,power-law index, and Reynolds number. At this juncture, it iscustomary to introduce the definitions of the aforementioneddimensionless groups, which will be used extensively in thediscussion of the results presented in the subsequent sections.

Reynolds Number.

The two critical Reynolds numbers denoting the onsets offlow separation and vortex shedding are denoted as Rec andRec, respectively. In the present work, the first transition (i.e.,Rec) was located by comparing the values of the dimensionlessstream function, vorticity, and velocity vector near the surfaceof the bluff body to check for the onset of flow separation.Typical values of both the dimensionless stream function[) ψ/U∞(2a)] and vorticity [) ω/U∞/(2a)] near the cylinderare on the order of 10-5-10-8.

The other critical Reynolds number, Rec, denotes the cessationof the steady-flow regime, and this point was located byexamining the lift coefficient-time history. For steady flow,the magnitude of the lift coefficient will gradually decrease withtime, ultimately becoming zero or very small. In this work, thecutoff value of the lift coefficient was on the order of 10-4.This transition was further substantiated by examining theasymmetry of the wake about the center line, y ) 0. Thus, inthe range Re < Rec, the flow is steady with two symmetricvortices, and for Re < Rec, the flow remains attached to thesurface of the bluff body.

Drag and Lift Coefficients. The drag and lift coefficientsare measures of the forces acting on the cylinder in the directionof flow (drag) and normal to flow (lift), respectively. They aredefined as

where FD and FL are the drag and lift forces per unit length ofthe cylinder, respectively.

Strouhal Number. The Strouhal number is a measure of thefrequency of vortex shedding and is defined as

where f is the frequency of vortex shedding and its value wasextracted from the lift coefficient-time behavior as a functionof the power-law index.

4. Numerical Solution Methodology

The present study was carried out using FLUENT (version6.3.26). The non-orthogonal “quadrilateral” cells of nonuniform

σ ) -pI + τ (3)

τ ) 2ηε(u) (4)

ε(u) ) 12

[∇U + (∇U)T] (5)

η ) m(I2

2 )(n-1)/2

(6)

Ux ) U∞;Uy ) 0 (7)

Ux ) 0;Uy ) 0 (8)

∂Ux

∂x) 0 and

∂Uy

∂x) 0 (9)

∂Ux

∂y) 0 and Uy ) 0 (10)

Re )FU∞

2-n(2a)n

m(11)

CD )FD

(1/2)FU∞2(2a)

(12)

CL )FL

(1/2)FU∞2(2b)

(13)

St ) (2a)fU∞

(14)

6652 Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010

grid spacing were generated using the commercial grid toolGAMBIT (version 2.3.16). At low Reynolds numbers, the flowis expected to be symmetric about the midplane, and the solutionwas sought only in a half-domain (y g 0) for identifying thepoint of flow separation. On the other hand, full-domaincomputations were carried out to delineate the transition fromthe steady symmetric to time-periodic vortex-shedding regime.The two-dimensional, laminar, segregated solver was used tosolve the incompressible flow on the collocated grid arrange-ment. The unsteady solver was used for the vortex-sheddingcase (Re > Rec), whereas the steady solver was used for locatingthe flow separation (Re ≈ Rec) case. The second-order upwindscheme was used to discretize the convective terms in themomentum equations. The semi-implicit method for pressure-linked equations (SIMPLE) scheme was used for solving thepressure-velocity coupling. Second-order time discretizationtogether with an implicit scheme was used for time integrationin this work. The “constant-density” and “non-Newtonianpower-law” viscosity models were used to input the physicalproperties of the fluid. This is, however, of no particularsignificance, as the final results are presented in a dimensionlessform. FLUENT solves the system of algebraic equations usingthe Gauss-Siedel (G-S) point-by-point iterative method inconjunction with the algebraic multigrid (AMG) method solver.The use of the AMG scheme can greatly reduce the number ofiterations (thereby accelerating convergence) and thus reducesthe CPU time required to obtain a converged solution, particu-larly when the model equations contain a large number ofcontrol volumes. The relative convergence criteria of 10-10 forthe continuity and x and y components of the momentumequations were prescribed in this work. Also, the solution wasconsidered to have converged when there was no change (atleast up to the fourth decimal place) in the total drag coefficientand the corresponding changes in the value of the lift coefficient(for the vortex-shedding case) were on the order of 10-5-10-6

for more than 1000 time steps, or when it showed more than10 constant periodic cycles in the time history of the lift anddrag coefficients.

5. Choice of Numerical Parameters

Undoubtedly, the accuracy and reliability of the numericalresults is strongly influenced by the choice of numericalparameters, namely, grid, domain size and time step, conver-gence criterion, and so on. Although all these issues were

addressed in detail in our previous studies,35,37 a short discussionof the salient features is included here.

5.1. Domain Size. It is well-known that the choices ofdomain size and grid size exert varying levels of influence onthe accuracy of numerical results. Because unconfined flow wassimulated in the current study, it was physically not possible toconsider infinite dimensions. Therefore, computations werecarried out by considering a domain of finite size, so it wasnecessary to obtain the size of the domain/grid that upon furtherincrement/refinement produced negligible changes in the nu-merical results.

Intuitively, it appears that the smaller the value of theReynolds number, the larger the value of the diameter of thecircular domain, D∞, required to minimize the boundary effects.This is obviously so because the boundary layer is very thickat low Reynolds numbers and thins with increasing Reynoldsnumber. Bearing this fact in mind, several values of D∞/(2a)ranging from 150 to 1200 were used in this study to choose anoptimal domain. A summary of representative results showingthe effect of domain size (value of D∞) on the value of the dragcoefficient is presented in Table 1 for a range of combinationsof the values of the power-law index (n), aspect ratio (E), andReynolds number (Re). An inspection of this table and of theother results not shown here reveals that, for Re < 5, the domainsize of D∞/(2a) ) 1200 is quite adequate for the results to befree from boundary effects. Similarly, the value of D∞/(2a) )300 is believed to be satisfactory for Re g 5. In fact, whereasthe drag values for n ) 0.3 and n ) 1.8 changed by less than1.5% upon increasing the value of D∞/(2a) from 1000 to 1200for Re ) 0.001 and E ) 0.2, the corresponding change for n )1 was about 2.9%. These small changes undoubtedly areaccompanied by a several-fold increase in CPU time. Whereasthe choice of D∞/(2a) ) 1200 for Re < 5 is in line with thestudy of Sivakumar et al.,37 clearly for Re g 5, the value ofD∞/(2a) ) 300 used here is larger than the value of 200 usedin previous studies.37,38 It is likely that a shorter domain (forRe < 5) would have sufficed, but only along with a much finergrid than that used here. Therefore, the choice of D∞ issomewhat linked to the choice of grid as well. Thus, D∞/(2a)) 1200 used here permits the use of a slightly coarse mesh,thereby leading to some reduction in CPU time. Moreover, thischoice also maintains consistency with our previous study.37

5.2. Grid Independence Study. Having fixed the domainsize, a grid independence study was carried out using three

Table 1. Effect of Domain Size on Drag Coefficient for Elliptical Cylinders of Different Aspect Ratios

CDP CD

domain size D∞/(2a) n ) 0.3 n ) 1.0 n ) 1.8 n ) 0.3 n ) 1.0 n ) 1.8

E ) 0.2, Re ) 0.001

800 22316.427 3321.2450 135.92970 23800.961 4130.8649 180.712451000 22316.414 3445.1481 132.18359 23800.960 3982.3009 175.731761200 22228.392 3215.5381 129.72411 23788.840 3868.5038 172.94004

E ) 0.2, Re ) 5

100 4.95522 3.21107 2.47850 5.32413 3.84600 3.25199200 4.95548 3.17827 2.44152 5.32439 3.80723 3.20353300 4.95547 3.16763 2.43007 5.32438 3.79465 3.18853

E ) 2, Re ) 1

450 13.54058 3.59730 1.85798 28.77424 11.23031 6.19543600 13.54057 3.58768 1.85279 28.77422 11.20092 6.178141200 13.54057 3.57284 1.84541 28.77423 11.15555 6.15358

E ) 5,Re ) 5

150 2.31150 1.02009 0.73787 8.29242 5.12557 3.741322200 2.31183 1.01478 0.73301 8.29299 5.10114 3.717230300 2.31199 1.00964 0.72840 8.29329 5.07750 3.694670

Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010 6653

nonuniform grids (G1, G2, and G3) for three values of power-law index (n) and a range of values of aspect ratio (E) andReynolds number (Re). A schematic representation of thenonuniform grid structure used in the present work is shown inFigure 1B. The grid was divided into four separate zones, andnonuniform grid distributions were employed. In the first zone(radius of 20a), the grid distribution was made fine around thecylinder to adequately capture the flow separation and thevortex-shedding phenomenon. The other three zones were madeup of relatively coarse cells, and the degree of coarsenessincreased toward the outer zone where the flow approached thefree streamflow and gradients were therefore expected to besmall as compared to that near the cylinder. Each grid wascharacterized in terms of the total number of cells (N) and thenumber of points on the surface of the cylinder (Np). A summaryof representative results is presented in Table 2. An inspectionof Table 2 clearly shows that a single grid would not besatisfactory for all values of E and/or of Reynolds number.Based on a detailed analysis of these results, the following gridswere used here to obtain the final results: For E < 1, grid G2for Re g 5 and grid G3 for Re < 5; for E ) 2, grid G2 for Reg 1; and for E ) 5, grid G1 for Re g 5. These choices are alsoin line with the grids used in our previous studies.37,38

5.3. Effect of Time Step. Evidently, the solution of time-dependent equations is required to delineate the value of Rec

denoting the onset of the time-dependent flow regime, andtherefore, the choice of an appropriate time step (∆T) alsoinfluences the accuracy of numerical results. The characteristictime (time period of oscillations) of the primary flow is givenby T ) (2a)/U∞, and the time step is defined as ∆T ) T/Ns,where Ns is the number of time steps per cycle. From Table 3,it can be seen that nearly identical results are obtained for thetwo vastly different time step values of 0.001 and 0.01 s. Inthis work, the choice of the time step was varied in this rangedepending on the level of convergence of the flow field. In thiswork, the time-marching numerical scheme was used with theinitial velocity profile corresponding to higher Reynolds numberand/or using a steady-flow field to initiate the calculations. In

experimental work, the flow can be destabilized by severalfactors including nonuniform flow conditions, vibrations dueto mechanical devices such as a pump or a compressor,irregularities of the boundary, surface roughness, and so on.Clearly, all of these factors are not present in numericalsimulations. On the other hand, even when the initial andboundary conditions are symmetric, the truncation, roundoff,and discretization errors can act as destabilizing factors thatultimately destabilize the flow, thereby leading to the vortex-shedding regime under appropriate conditions of the pertinentgoverning factors such as power-law index, aspect ratio, andReynolds number. Naturally, this approach requires a signifi-cantly large time for these errors to grow to the extent that willultimately lead to the onset of time-dependent periodic-flowregime. To obviate this difficulty, the flow can be perturbedartificially, which accelerates the establishment of the time-dependent flow regime, without, of course, influencing theultimate transitional values of the Reynolds and Strouhalnumber. This technique was used by Braza et al.40 andsubsequently by others9 relating to the case of a circular cylinder.Although it is relatively straightforward to perturb the flow byrotating a circular cylinder for a short period, the relevance ofthis method is less clear for an elliptical cylinder. Therefore, inthe present study, no external stimulus was used to acceleratethe establishment of the fully periodic-flow regime. In otherwords, time-dependent calculations were continued for a longtime until either the flow became steady or fully periodic intime. The time step was progressively increased from the initialvalue of 0.001 to 0.01 s as the solution approached the desiredlevel of convergence. The simulations were carried out untilthe time history of lift coefficients either died out or reachedsteady periodic oscillations. The critical values of the parametersReynolds number, Rec; Strouhal number, Stc; and drag coef-ficients at the critical conditions are presented in the next section.

All computations were performed on Quad Core PCs (2.83GHz, 3 GB RAM) operating in Windows. A typical run on a

Table 2. Effect of Mesh on Drag Coefficient for Elliptical Cylinders of Different Aspect Ratios

CDP CD

grid no. of cells (N) Np n ) 0.3 n ) 1.0 n ) 1.8 n ) 0.3 n ) 1.0 n ) 1.8

E ) 0.2, Re ) 0.001

G1 32500 200 22308.543 3313.7746 132.85307 23795.345 3966.4860 175.78489G2 65000 400 22323.601 3324.543 131.81055 23807.693 3989.2157 175.58578G3 48750 300 22316.414 3321.245 132.18359 23800.960 3982.3009 175.73176

E ) 0.2, Re ) 5

G1 51000 200 4.95330 3.18430 2.45524 5.31385 3.80772 3.20475G2 55500 400 4.87664 3.09505 2.35520 5.32471 3.80072 3.19839G3 82500 300 4.95548 3.17827 2.44152 5.32439 3.80723 3.20353

E ) 2, Re ) 1

G1 20500 200 13.50197 3.58208 1.85469 28.73837 11.18656 6.18016G2 27750 300 13.87037 3.69345 1.89696 28.81323 11.20777 6.17723G3 30750 300 13.54057 3.58768 1.85279 28.77422 11.20092 6.17814

E ) 5, Re ) 5

G1 50000 200 2.31549 1.01852 0.73591 8.292613 5.10219 3.71884G2 76000 300 2.29304 0.99612 0.71352 8.290613 5.09931 3.71563G3 82500 400 2.31183 1.00964 0.73301 8.292990 5.10114 3.71723

Table 3. Effects of Time Step on Drag Coefficient, Strouhal Number, and Lift Coefficient

E ) 0.2 n ) 0.3 n ) 1.0 n ) 1.8

∆T CD St CL CD St CL CD St CL

0.001 2.1351 0.1760 (0.3403 1.7681 0.1250 (0.0648 1.9851 0.1149 (0.08260.01 2.1353 0.1759 (0.3402 1.768 0.1249 (0.0648 1.9853 0.1149 (0.0826

6654 Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010

grid with 48750 cells for n ) 0.3 and E ) 0.2 required around14 days of computing to obtain 10 stable cycles of vortexshedding.

6. Results and Discussion

6.1. Validation of Results. Prior to presenting the newresults on the delineation of the critical Reynolds numbers,it is worthwhile to validate the choices of numericalparameters and numerical methodology employed here. Overthe years, extensive numerical results have been reported inthe literature on the values of drag coefficients for the steady2-D flow of Newtonian fluids over circular and ellipticalcylinders. Suffice to say here that the present values for acircular cylinder are within 3% of that reported previouslyby several authors34,37,41-45 for Re ) 1, 5, and 40. Table 4shows comparisons between the present and literature valuesfor a range of values of E and Re. Excellent correspondenceis seen to exist between the present and literature values.Furthermore, for E ) 5, the present results (not shown here)are within 1% of those reported by Dennis and Young17 andD’Alessio and Dennis16 for Re ) 5 and 40. Table 5 presentsa similar comparison for the steady flow of power-law fluidsat Re ) 5 and 40 where, once again, good match is seen toexist. Also, the present numerical solution procedure wasvalidated by comparing the critical Reynolds numbers inNewtonian fluids for the onset of wake formation (Table 6)and the onset of vortex shedding (Table 7) for ellipticalcylinders of different aspect ratios. Once again, the agreementwith the literature values is seen to be good. Such a close

correspondence between the present and literature values overwide ranges of conditions inspires confidence and lendscredibility to the reliability and accuracy of the resultsreported herein. Suffice it to add here that the order ofdifferences seen in Tables 4-7 are not at all uncommon innumerical studies and are often attributed to differences ingrid, domain, and solution methods used in different studies.51

6.2. Onset of Wake Formation. The appearance of wakeswas ascertained here by the inspection of the computedvelocity vector field at the critical and nearby values of theReynolds number. When the flow remains attached to thesurface of the cylinder, velocity vectors in the proximity ofthe cylinder all point in the downstream direction; the onsetof wake formation manifests itself by the fact that somevelocity vectors near the cylinder point in the upstreamdirection and form a close recirculating flow region. Repre-sentative results on the streamline patterns and velocityvectors in the vicinity of cylinder for n ) 0.3 and n ) 1.6 attwo closeby values of Reynolds number that bracket thetransition point are shown in Figure 2A,B. It can be seen inthis figure that, at the smaller value of the Reynolds number(left column), no wake forms, and all velocity vectors nearthe cylinder (not shown here) point in the downstreamdirection, whereas at the higher value of the Reynolds number(right column), some of the velocity vectors are oriented inthe direction opposite to the main flow, which denotes flowseparation and the formation of a wake. For instance, inFigure 2A, for n ) 0.3 and E ) 0.2, the flow remains attachedto the surface of the cylinder at Re ) 1.4, whereas at Re )1.5, separation has occurred in the rear of the cylinder. Thus,for n ) 0.3, the critical Reynolds number (Rec) lies in therange 1.4 e Re e 1.5 for an elliptical cylinder of aspectratio 0.2. This approach was used to ascertain the intervalfor flow separation for the other values of the power-lawindex and aspect ratio. The resulting values of the criticalReynolds number for E ) 0.5, 2, and 5 are in the intervals6.4 e Re e 6.6, 32 e Re e 33, and 98 e Re e 99,respectively. As expected, the critical Reynolds numberincreases with increasing degree of streamlining of the bluffbody (i.e., with increasing value of E). The ranges of criticalReynolds numbers for all values of power-law index, n, andaspect ratio, E, considered herein are summarized in Table8 and shown in Figure 3. In this figure, the top and bottomlines represent the most (E ) 5) and least (E ) 0.2)streamlined shapes, respectively. As expected, bluffness ofthe cylinder causes wake formation at lower values ofReynolds number for both shear-thinning and shear-thicken-ing fluids. The flow separation that causes wake formationbehind the cylinder occurs at higher Reynolds number inshear-thinning fluids and continues to decrease as the power-law index increases for Newtonian and shear-thickening

Table 4. Steady Newtonian Flow over an Elliptical Cylinder

CD

source Re ) 0.01 Re ) 5

E ) 0.2

present value 400.6309 3.7820Sivakumar et al.37 404.5256 3.7900Dennis and Young17 - 3.8540D’Alessio and Dennis16 - 3.8620

E ) 1

present value - 3.9234Juncu45 - 3.8930

E ) 5

present value 471.1033 5.0710Sivakumar et al37 471.7567 5.0715

Table 5. Steady Power-Law Fluid Flow over an Elliptical Cylinder

Re ) 5 Re ) 40

CD CD

source n ) 0.6 n ) 1.8 n ) 0.6 n ) 1.8

E ) 2

present value 5.1008 3.3448 1.3209 1.6507Sivakumar et al.37 5.1173 3.3586 1.3214 1.6523

E ) 5

present value 6.5487 3.7180 1.3450 1.7061Sivakumar et al.37 6.5021 3.6928 1.3470 1.7025

Table 6. Critical Reynolds Number for Wake Formation forNewtonian Flow over an Elliptical Cylinder

source E Rec

present value 0.2 0.85Stack and Bravo25 0.84present value 0.5 3.0Stack and Bravo25 2.85

Table 7. Critical Reynolds Number for Onset of Vortex Sheddingfor Newtonian Flow over an Elliptical Cylinder

source E Rec Stc

present value 0.5 40 0.1181Jackson23 0.5 35.704 0.1332Johnson et al.20 0.5 42.5-45 -present value 1 48 0.1149Sivakumar et.al.9 1 47 0.1172Jackson23 1 45.403 0.1362Norberg46,47 1 47.5((0.5) 0.1220Williamson48 1 47.9 0.1168Kumar and Mittal49 1 47.336 0.1168Morzynski et al.50 1 47.0 0.1320present value 2 78 0.1266Jackson23 2 76.794 0.1464

Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010 6655

fluids. This form of dependence of the critical Reynoldsnumber on the power-law index for elliptical cylinders ofaspect ratio E ) 2 and 5 is similar to that seen for a circularcylinder. For a fixed shape of the cylinder (i.e., for a given

value of the aspect ratio), the dependence of the criticalReynolds number on the flow behavior index can be explainedas follows: A fluid element experiences the maximum rateof deformation near the bluff body, and it decays as one

Figure 2. Streamline patterns near elliptical cylinders with aspect ratios of 0.2, 0.5, 2, and 5 at low Reynolds number values for (A) n ) 0.3 and (B) n )1.6.

6656 Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010

moves away from the cylinder. In other words, for a shear-thinning fluid (n < 1), the effective viscosity is minimumclose to the surface of the bluff body and increases rapidlyaway from the cylinder. Thus, the low-viscosity region closeto the surface is surrounded by an envelope of fluid ofincreasing viscosity, which tends to suppress the propensityfor flow separation. This is analogous to imposing a solidboundary to constrict the flow. Of course, in a shear-thickening fluid (n > 1), the effective viscosity will graduallydecrease away from the bluff body. This explains qualitativelythe fact that the flow remains attached to the surface of thecylinder up to higher values of Reynolds number in shear-thinning fluids and flow separation is advanced to a lowervalue of the Reynolds number in shear-thickening fluids. Ofcourse, the rate of deformation is strongly influenced by theshape of the body, especially for the extreme case of E )0.2, and the weak maximum seen in Figure 3 is clearly dueto the complex interplay between the shape (value of E) andthe highly shear-thinning nature of the fluid (n ≈ 0.2-0.3)wherein the velocity field decays rather quickly.

6.3. Onset of Vortex Shedding. Once wake formation hasset in, the size of this region grows as the Reynolds numberis progressively increased. The flow is steady and symmetricabout the midplane, however. Finally, a point (value of Redepending on the value of n) is reached where the standingvortices lose symmetry and begin to flap. Only upon a slightfurther increase in the value of the Reynolds number are thestanding vortices convected by the bulk flow in the down-stream direction. In the flapping zone, simulations were runfor a sufficiently long time to ensure that the vortices flaponly, without exhibiting vortex shedding. The transition fromthe flapping to shedding of vortices occurs over a very narrowrange of Reynolds number, and this range depends primarilyon the values of the aspect ratio of the cylinder and thepower-law index. Also, this range of Reynolds number whereonly flapping of vortices occurs was found to be a bit widerin the case of elliptical cylinders with aspect ratios greaterthan 1. The onset of vortex shedding was judged byexamining the values of the vorticity and stream functioncontours near the surface of the cylinder along with theevolution of the lift coefficient with time. The vorticitycontours in the vicinity of elliptical cylinders of aspect ratio0.2, 0.5, 2, and 5 for two values of the Reynolds numberbracketing the transition and for extreme values of the power-law index 0.3 and 1.8 are shown in Figure 4A and B,respectively. In these figures, depending on the values of theaspect ratio and power-law index, vortices are seen to beeither standing or flapping at the lower value of the Reynoldsnumber (left column), whereas at the higher Reynolds number(right column), shedding of vortices has initiated. Forinstance, in Figure 4A, for n ) 0.3 and E ) 0.2, the flow isseen to be asymmetric because of the flapping of the vorticesat Re ) 32, whereas at Re ) 33, vortex shedding is clearlyseen to have started. Because the flow is asymmetric at Re) 32, the cylinder experiences a lift force and the liftcoefficient oscillates between +0.019 and -0.019 with aconstant frequency. The lift coefficient at Re ) 32 is nearlysame as that at Re ) 33. However, for the cases where theflow is steady and symmetric at lower Reynolds number (inFigure 4B for n ) 1.8 and E ) 0.2), the lift coefficientultimately approaches zero as the time progresses. Therefore,for the elliptical cylinder of aspect ratio 0.2, the criticalReynolds number for power-law index of 0.3 lies in the range32 e Re e 33. Qualitatively, the same procedure is followedto infer the value of Rec for the other values of the power-law index and the aspect ratio. Figure 5 shows the dependenceof the critical Reynolds number (Rec) on the power-law indexand aspect ratio. It can be seen that the critical Reynoldsnumber is higher for shear-thinning fluids than for shear-thickening fluids only for n g 0.6. As the power-law indexincreases from 0.3 to 1.0, the critical Reynolds number goesthrough a maximum value at about n ) 0.6, which is similarto that seen for a circular cylinder.9 On the other hand, forshear-thickening fluids, the decrease in the critical Reynoldsnumber is rather steep as the fluid behavior transits fromNewtonian (n ) 1) to shear-thickening (n > 1). The complexdependence of the critical Reynolds number on the flowbehavior index seen in Figure 5 stems from the interactionbetween the two nonlinear terms (inertial and viscous) presentin the momentum equations. These two nonlinear terms scaledifferently with velocity. The viscous term scales as U∞

n

whereas the inertial term scales as U∞2. Thus, at low Reynolds

numbers, the viscous term diminishes with decreasing valueof a power-law index for shear-thinning fluids, but it increases

Table 8. Effect of Power-Law Index on the Critical ReynoldsNumber (Rec) for Flow with and without Wake Formation

E ) 0.2 E ) 0.5

n Re (no wake) Rec (wake) n Re (no wake) Rec (wake)

0.3 1.4 1.5 0.3 6.4 6.60.4 2.1 2.2 0.4 6.6 6.70.6 2.3 2.35 0.6 5.8 5.90.8 1.5 1.7 0.8 4.5 4.61 0.8 0.85 1 2.9 31.2 0.3 0.4 1.2 1.9 21.4 0.6 0.2 1.4 1.1 1.21.6 0.006 0.01 1.6 0.4 0.6

1.8 0.2 0.3

E ) 2 E ) 5

n Re (no wake) Rec (wake) n Re (no wake) Rec (wake)

0.3 32 33 0.3 98 990.4 30 31 0.4 89 900.6 24 25 0.6 71 720.8 19 20 0.8 55 561 14 15 1 37 381.2 10 11 1.2 26 271.4 6 6.4 1.4 18 191.6 4 4.5 1.6 11 121.8 2 3 1.8 7 8

Figure 3. Effects of power-law index and aspect ratio on critical Reynoldsnumber for onset of wake formation (results for E ) 1 from ref 9).

Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010 6657

with velocity in shear-thickening fluids. Also, the viscosityof a shear-thinning fluid becomes very large as the shear ratedecreases, and it thus tends to an infinite value far from thecylinder where the shear rate tends to zero. Therefore, theviscous effects always dominate the flow characteristics even

far from the cylinder under these conditions. Conversely, theviscosity of shear-thickening fluids is maximum near theobstacle and gradually decreases away from the object. Fora fixed value of the power-law index, as the aspect ratio ofthe cylinder is increased from 0.2 to 5, shedding is delayed

Figure 4. Vorticity contours near elliptical cylinders with aspect ratios of 0.2, 0.5, 2, and 5 for (A) n ) 0.3 and (B) n ) 1.8.

6658 Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010

to higher Reynolds numbers. This is due to the increasingstreamlining of the body, which delays the boundary layerseparation.

Figure 6 shows the variation of the critical Strouhal number(Stc) at the corresponding critical Reynolds number withpower-law index (0.3 e n e 1.8) and aspect ratio (0.2 e Ee 5). An increase in power-law index results in a decreasein the critical Strouhal number, and the trend is seen to besimilar for all values of aspect ratio. This can be explainedas follows: All else being equal, the effective viscosity inthe vicinity of the cylinder gradually increases with increasingvalue of the power-law index, which tends to suppress vortexshedding. The corresponding values of the time-average dragcoefficient are shown in Figure 7. It can be seen that thetime-average drag coefficient increases with increasingpower-law index (n), which is in line with the trend seen inthe steady-flow regime.37 For aspect ratios less than 1, thedrag coefficient decreases as the value of the power-law indexis increased from 0.3 to 0.4, and with a further increase inthe power-law index from 0.4 to 1.8, the drag coefficientcontinuously increases. For aspect ratios greater than 1, as

the power-law index is increased from 0.3 to 1.8, the dragcoefficient rises continuously. Thus, highly shear-thinningfluids exhibit different behavior than mildly shear-thinningand shear-thickening fluids. In view of these results, it isfair to say that the literature values of drag coefficient andNusselt number based on the assumption of steady flow forRe > 30, E ) 0.2-5 and/or n g 0.2 are likely to be ratherinaccurate.37,38

7. Conclusions

Extensive numerical results on the critical Reynolds numbersdenoting wake formation and the onset of vortex shedding forthe flow of Newtonian and power-law fluids have been obtainedover wide ranges of conditions: 0.2 e E e 5 and 0.3 e n e1.8. For elliptical cylinders with aspect ratios varying from 0.2to 5, wake formation is seen to be delayed as the fluid behaviorchanges from shear-thickening to Newtonian and finally toshear-thinning. In the case of shear-thinning fluids, for aspectratios less than 1, the critical Reynolds number goes through amaximum value as the fluid nature changes from highly shear-thinning to mildly shear-thinning. For shear-thickening fluids,the critical Reynolds number decreases monotonically withpower-law index for all values of aspect ratio of ellipticalcylinders considered here. Furthermore, the critical Reynoldsnumber curve for the onset of vortex shedding for shear-thinningfluids also goes through a maximum at n ) 0.6, whereas itdecreases continually with power-law index for shear-thickeningfluids for all values of aspect ratio. It is also seen that,irrespective of the value of aspect ratio (0.2 e E e 5), vortexshedding occurs at lower Reynolds number in shear-thickeningfluids than in Newtonian fluids. Therefore, shear-thinningbehavior seems to have a stabilizing influence, and shear-thickening behavior tends to induce instability in the flow atlower and lower values of Reynolds number. In addition, theonsets of wake formation and vortex shedding in both shear-thinning and shear-thickening fluids are seen to be delayed tohigher Reynolds number as the aspect ratio of the ellipticalcylinder increases from 0.2 to 5 because of the increasingstreamlining of the shape of the body.

Figure 5. Effects of power-law index and aspect ratio on critical Reynoldsnumber for the onset of vortex shedding (results for E ) 1 from ref 9).

Figure 6. Effects of power-law index and aspect ratio on critical Strouhalnumber at the onset of vortex shedding (results for E ) 1 from ref 9).

Figure 7. Effects of power-law index and aspect ratio on time-average dragcoefficient at the onset of vortex shedding (results for E ) 1 from ref 9).

Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010 6659

Notationa ) semiaxis of the elliptical cylinder normal to the direction of

flow, m

b ) semiaxis of the elliptical cylinder along the direction offlow, m

CD ) drag coefficient [ )FD

(1/2)FU∞2(2a)]

CjD ) time-average drag coefficient

CDF ) friction component of drag coefficient [ )FDF

(1/2)FU∞2(2a)]

CDP ) pressure component of drag coefficient [ )FDP

(1/2)FU∞2(2a)]

CL ) lift coefficient [ )FL

(1/2)FU∞2(2b)]

D∞ ) diameter of the outer boundary of the circular domain, m

E ) aspect ratio () b/a)

f ) frequency of vortex shedding, s-1

FD ) drag force per unit length of cylinder, N ·m-1

FDP ) pressure force per unit length of cylinder, N ·m-1

FL ) lift force per unit length of cylinder, N ·m-1

I2 ) second invariant of the rate of the strain tensor, s-2

m ) power-law consistency index, Pa · sn

n ) power-law flow behavior index

Np ) number of grid points on the surface of the cylinder

p ) pressure, Pa

Re ) Reynolds number [ )FU∞

2-n(2a)n

m ]St ) Strouhal number [) (2a)f/U∞]

Ux, Uy ) x and y components of the velocity, m · s-1

U∞ ) free stream velocity, m · s-1

x ) stream-wise coordinates, m

y ) transverse coordinates, m

Greek Symbols

∆t ) time step, s

ε ) component of the rate of the strain tensor, s-1

η ) viscosity, Pa · sF ) density of the fluid, kg ·m-3

τ ) extra stress, Pa

φ ) scalar variable (Ux,Uy), m · s-1

ψ ) stream function, m2 · s-1

ω ) vorticity, s-1

Superscript/Subscript

c ) critical

Literature Cited

(1) Zdravkovich, M. M. Flow Around Circular Cylinders, Vol. 1:Fundamentals; Oxford University Press: New York, 1997.

(2) Zdravkovich, M. M. Flow Around Circular Cylinders, Vol. 2:Applications; Oxford University Press: New York, 2003.

(3) Coutanceau, M.; Defaye, J. R. Circular cylinder wake configuration:A flow visualization survey. Appl. Mech. ReV. 1991, 44, 255–305.

(4) Raynor, P. C. Flow field and drag for elliptical filter fibers. AerosolSci. Technol. 2002, 36, 1118–1127.

(5) Raynor, P. C. Single-fiber interception efficiency for elliptical fibers.Aerosol Sci. Technol. 2008, 42, 357–368.

(6) Wang, J.; Pui, D. Y. H. Filtration of aerosol particles by ellipticalfibers: A numerical study. J Nanopart. Res. 2009, 11, 185–196.

(7) Janusz, K.; Chen, L. H.; Russell, H.; Patterson, R.; Ackerman, E.Contributions to the impedance cardiogram waveform. Ann. Biomed. Eng.1986, 14, 67–80.

(8) Chhabra, R. P.; Richardson, J. F. Non-Newtonian Flow and AppliedRheology: Engineering Applications, 2nd ed.; Butterworth-Heinemann:Oxford, U.K., 2008.

(9) Sivakumar, P.; Bharti, R. P.; Chhabra, R. P. Effect of power-lawindex on critical parameters for power-law fluid flow across an unconfinedcircular cylinder. Chem. Eng. Sci. 2006, 61, 6035–6046.

(10) Hasimoto, H. On the flow of a viscous fluid past an inclined ellipticcylinder at small Reynolds numbers. J. Phys. Soc. Jpn. 1958, 8, 653–661.

(11) Imai, I. A new method of solving Oseen’s equations and itsapplications to the flow past an inclined elliptic cylinder. Proc. R. Soc.London A 1954, 224, 141–160.

(12) Shintani, K.; Umemura, A.; Takano, A. Low-Reynolds-number flowpast an elliptic cylinder. J. Fluid Mech. 1983, 136, 227–289.

(13) Tomotika, S.; Aoi, T. The steady flow of a viscous fluid past anelliptic cylinder and a flat plate at small Reynolds number. Q. J. Mech.Appl. Math. 1953, 6, 290–312.

(14) Epstein, N.; Masliyah, J. H. Creeping flow through clusters ofspheroids and elliptical cylinders. Chem. Eng. J. 1972, 3, 169–175.

(15) Meller, N. A. Viscous flow past an elliptic cylinder. Comput. Math.Math. Phys. 1978, 18, 138–149.

(16) D’Alessio, S. J. D.; Dennis, S. C. R. A vorticity model for viscousflow past a cylinder. Comput. Fluids 1994, 23, 279–293.

(17) Dennis, S. C. R.; Young, P. J. S. Steady flow past an elliptic cylinderinclined to the stream. J. Eng. Math. 2003, 47, 101–120.

(18) Lugt, H. J.; Haussling, H. J. Laminar flow past an abruptlyaccelerated elliptic cylinder at 450 incidence. J. Fluid Mech. 1974, 65, 711–734.

(19) Patel, V. A. Flow around the impulsively started elliptic cylinderat various angles of attack. Comput. Fluids 1981, 9, 435–462.

(20) Johnson, S. A.; Thompson, M. C.; Hourigan, K. Flow past ellipticalcylinders at low Reynolds numbers. In Proceedings of the 14th AustralasianFluid Mechanics Conference; University of Adelaide: Adelaide, Australia,2001; pp 343-346.

(21) Ahmad, E. H.; Badr, H. M. Mixed convection from an elliptic tubeplaced in a fluctuating free stream. Int. J. Eng. Sci. 2001, 39, 669–693.

(22) Khan, W. A.; Culham, J. R.; Yovanovich, M. M. Fluid flow aroundand heat transfer from elliptical cylinders: Analytical approach. J. Ther-mophys. Heat Transfer 2005, 19, 178–185.

(23) Jackson, C. P. A finite-element study of the onset of vortex sheddingin flow past variously shaped bodies. J. Fluid Mech. 1987, 182, 23–45.

(24) Faruquee, Z.; Ting, D. S-K.; Fartaj, A.; Barron, R. M.; Carriveau,R. The effects of axis ratio on laminar fluid flow around an elliptical cylinder.Int. J Heat Fluid Flow 2007, 28, 1178–1189.

(25) Stack, D.; Bravo, H. R. Flow separation behind ellipses at Reynoldsnumbers less than 10. Appl. Math. Model. 2009, 33, 1633–1643.

(26) Chhabra, R. P.; Soares, A. A.; Ferreira, J. M. Steady non-Newtonianflow past a circular cylinder: A numerical study. Acta Mech. 2004, 172,1–16.

(27) D’Alessio, S. J. D.; Pascal, J. P. Steady flow of power-law fluidspast a cylinder. Acta Mech. 1996, 117, 87–100.

(28) Soares, A. A.; Ferreira, J. M.; Chhabra, R. P. Flow and forcedconvection heat transfer in cross flow of non-Newtonian fluids over a circularcylinder. Ind. Eng. Chem. Res. 2005, 44, 5815–5827.

(29) Bharti, R. P.; Chhabra, R. P.; Eswaran, V. Steady flow of powerlaw fluids across a circular cylinder. Can. J. Chem. Eng. 2006, 84, 406–421.

(30) Tanner, R. I. Stokes paradox for power-law fluids around a cylinder.J. Non-Newtonian Fluid Mech. 1993, 50, 217–224.

(31) Bharti, R. P.; Chhabra, R. P.; Eswaran, V. Two-dimensional steadyPoiseuille flow of power-law fluids across a circular cylinder in a planeconfined channel: Wall effect and drag coefficient. Ind. Eng. Chem. Res.2007, 46, 3820–3840.

(32) Bharti, R. P.; Chhabra, R. P.; Eswaran, V. Effect of blockage onheat transfer from a cylinder to power-law liquids. Chem. Eng. Sci. 2007,62, 4729–4741.

(33) Srinivas, A. T.; Bharti, R. P.; Chhabra, R. P. Mixed convectionheat transfer from a cylinder in power-law fluids: Effect of aiding buoyancy.Ind. Eng. Chem. Res. 2009, 48, 9735–9754.

(34) Soares, A. A.; Anacleto, J.; Caramelo, L.; Ferreira, J. M.; Chhabra,R. P. Mixed convection from a circular cylinder to power-law fluids. Ind.Eng. Chem. Res. 2009, 48, 8219–8231.

(35) Patnana, V. K.; Bharti, R. P.; Chhabra, R. P. Two-dimensionalunsteady flow of power-law fluids over a cylinder. Chem. Eng. Sci. 2009,64, 2978–2999.

6660 Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010

(36) Patnana, V. K.; Bharti, R. P.; Chhabra, R. P. Two-dimensionalunsteady forced convection heat transfer in power-law fluids from a cylinder.Int. J. Heat Mass Transfer, in press.

(37) Sivakumar, P.; Bharti, R. P.; Chhabra, R. P. Steady flow of power-law fluids across an unconfined elliptical cylinder. Chem. Eng. Sci. 2007,62, 1682–1702.

(38) Bharti, R. P.; Sivakumar, P.; Chhabra, R. P. Forced convectionheat transfer from an elliptical cylinder to power-law fluids. Int. J. HeatMass Transfer 2008, 51, 1838–1853.

(39) Bird, R. B.; Stewart, W. E.; Lightfoot, E. N. Transport Phenomena;Wiley: New York, 2002.

(40) Braza, M.; Chassaing, P.; Ha Minh, H. Numerical study andphysical analysis of the pressure and velocity fields in the near wake of acircular cylinder. J. Fluid Mech. 1986, 165, 79–130.

(41) Dennis, S. C. R.; Chang, G. Z. Numerical solutions for steady flowpast a circular cylinder at Reynolds numbers up to 100. J. Fluid Mech.1970, 42, 471–489.

(42) Fornberg, B. A numerical study of steady viscous flow past acircular cylinder. J. Fluid Mech. 1980, 98, 819–855.

(43) Park, J.; Kwon, K.; Choi, H. Numerical solutions of flow past acircular cylinder at Reynolds numbers up to 160. KSME Int. J. 1998, 12,1200–1205.

(44) Chakraborty, J.; Verma, N.; Chhabra, R. P. Wall effects in the flowpast a circular cylinder in a plane channel: A numerical study. Chem. Eng.Process. 2004, 43, 1529–1537.

(45) Juncu, G. A numerical study of momentum and forced convectionheat transfer around two tandem circular cylinders at low Reynolds numbers.Part I: Momentum transfer. Int. J. Heat Mass Transfer 2007, 50, 3788–3798.

(46) Norberg, C. An experimental investigation of the flow around a circularcylinder: Influence of aspect ratio. J. Fluid Mech. 1994, 258, 287–316.

(47) Norberg, C. Flow around a circular cylinder: Aspects of fluctuatinglift. J. Fluids Struct. 2001, 15, 459–469.

(48) Williamson, C. H. K. Oblique and parallel modes of vortex sheddingin the wake of a circular cylinder at low Reynolds numbers. J. Fluid Mech.1989, 206, 579–627.

(49) Kumar, B.; Mittal, S. Effect of blockage on critical parameters forflow past a circular cylinder. Int. J. Numer. Methods Fluids 2006, 50, 987–1001.

(50) Morzynski, M.; Afanasiev, K.; Thiele, F. Solution of the eigen valueproblems resulting from global non-parallel flow stability analysis. Comput.Methods Appl. Mech. Eng. 1999, 169, 167–176.

(51) Roache, P. J. Perspective: A method for uniform reporting of gridrefinement studies. J. Fluids Eng. 1994, 116, 405–413.

ReceiVed for reView February 2, 2010ReVised manuscript receiVed May 12, 2010

Accepted June 2, 2010

IE100251W

Ind. Eng. Chem. Res., Vol. 49, No. 14, 2010 6661