poole_16

11
Flow produced in a conical container by a rotating endwall M.P. Escudier * , J. O’Leary, R.J. Poole Department of Engineering, University of Liverpool, Brownlow Street, Liverpool L69 3GH, United Kingdom Received 5 October 2006; received in revised form 16 April 2007; accepted 18 April 2007 Available online 13 June 2007 Abstract Numerical calculations have been carried out for flow in a truncated cone generated by rotation of one endwall. For both convergent (radius increasing with approach to the rotating endwall) and divergent geometries, vortex breakdown is suppressed beyond a certain angle of inclination of the sidewall. At the same time Moffat eddies of increasing strength and extent appear in the corner between the sidewall and the non-rotating endwall. For the divergent geometry, a zone of recirculation appears on the sidewall and eventually merges with the Moffat eddies. The flow phenomena identified from streamline patterns are consistent with the calculated variation of pressure around the periphery of the computational domain. Ó 2007 Elsevier Inc. All rights reserved. Keywords: Vortex breakdown; Conical container; Rotating endwall 1. Introduction Rotation of one endwall of a closed cylinder completely filled with a fluid produces a wide range of flows from unre- markable steady swirling recirculation to a flow with up to three separate zones of embedded recirculation (vortex breakdowns) centred on the cylinder axis, to periodic time-varying flows and eventually to a fully turbulent flow. The flow structure is determined by two global parameters: the rotational Reynolds number, Re = XR 2 /t, and the aspect ratio of the cylinder, H/R. X is the angular velocity of the rotating endwall and R its radius, H is the cylinder height and t is the kinematic viscosity of the fluid. Vogel’s (1968) flow-visualisation experiments revealed the presence of a single region of recirculation for 1000 < Re < 2870 and 1.39 < H/R < 2.12 whereas a more complete picture, also based upon flow visualisation, was given by Escudier (1984) who extended the upper limits of the parameter ranges to Re 3480 and H/R 3.65. Since that time a number of studies have involved the investigation of flows in closely related geometries. Pereira and Sousa (1999) modified the experimental configuration to include a coni- cal rotating endwall, Fujimura et al. (2004) investigated the flow generated by rotation of both endwalls, while Spohn et al. (1998) carried out experiments on flows with a free surface. Bu ¨ hler (1985) reported on experiments for the flow in the gap between two concentric spheres, the outer of which was rotating, and on the flow in a conical container with curved endwalls, one of which was rotating (Bu ¨ hler (1994)). Given the combination of these experiments with excep- tionally well defined, quite simple boundary conditions and an extraordinarily complex flowfield, it is unsurprising that there have also been numerous numerical studies of the ‘container with endwall rotation’ problem. Of note are the early works of Lugt and Haussling (1982), which was limited to calculating the single recirculation bubble, and the detailed simulations of Lopez (1990) who was the first to calculate the full extent of the flows observed by Escu- dier. Only recently have the numerical investigations predicted phenomena not previously observed experimen- tally. For example, Tsitverblit and Kit (1996) and Mullin et al. (1998) investigated numerically the inclusion of a cen- tral circular cylinder and found it to have a relatively small 0142-727X/$ - see front matter Ó 2007 Elsevier Inc. All rights reserved. doi:10.1016/j.ijheatfluidflow.2007.04.018 * Corresponding author. Fax: +44 151 7944848. E-mail address: [email protected] (M.P. Escudier). www.elsevier.com/locate/ijhff Available online at www.sciencedirect.com International Journal of Heat and Fluid Flow 28 (2007) 1418–1428

Upload: ryan-gonzales

Post on 30-Sep-2015

214 views

Category:

Documents


0 download

DESCRIPTION

Heat Transfer

TRANSCRIPT

  • on

    O

    Bro

    d fone

    ateergiesetrom

    2007 Elsevier Inc. All rights reserved.

    The ow structure is determined by two global parameters:the rotational Reynolds number, Re = XR2/t, and the

    ranges to Re 3480 and H/R 3.65. Since that time anumber of studies have involved the investigation of ows

    with curved endwalls, one of which was rotating (Buhler(1994)).

    to calculate the full extent of the ows observed by Escu-dier. Only recently have the numerical investigationspredicted phenomena not previously observed experimen-tally. For example, Tsitverblit and Kit (1996) and Mullinet al. (1998) investigated numerically the inclusion of a cen-tral circular cylinder and found it to have a relatively small

    * Corresponding author. Fax: +44 151 7944848.E-mail address: [email protected] (M.P. Escudier).

    Available online at www.sciencedirect.com

    International Journal of Heat and Fluidaspect ratio of the cylinder, H/R. X is the angular velocityof the rotating endwall and R its radius, H is the cylinderheight and t is the kinematic viscosity of the uid. Vogels(1968) ow-visualisation experiments revealed the presenceof a single region of recirculation for 1000 < Re < 2870 and1.39 < H/R < 2.12 whereas a more complete picture, alsobased upon ow visualisation, was given by Escudier(1984) who extended the upper limits of the parameter

    Given the combination of these experiments with excep-tionally well dened, quite simple boundary conditions andan extraordinarily complex oweld, it is unsurprising thatthere have also been numerous numerical studies of thecontainer with endwall rotation problem. Of note arethe early works of Lugt and Haussling (1982), which waslimited to calculating the single recirculation bubble, andthe detailed simulations of Lopez (1990) who was the rstKeywords: Vortex breakdown; Conical container; Rotating endwall

    1. Introduction

    Rotation of one endwall of a closed cylinder completelylled with a uid produces a wide range of ows from unre-markable steady swirling recirculation to a ow with up tothree separate zones of embedded recirculation (vortexbreakdowns) centred on the cylinder axis, to periodictime-varying ows and eventually to a fully turbulent ow.

    in closely related geometries. Pereira and Sousa (1999)modied the experimental conguration to include a coni-cal rotating endwall, Fujimura et al. (2004) investigated theow generated by rotation of both endwalls, while Spohnet al. (1998) carried out experiments on ows with a freesurface. Buhler (1985) reported on experiments for the owin the gap between two concentric spheres, the outer ofwhich was rotating, and on the ow in a conical containerFlow produced in a conical c

    M.P. Escudier *, J.

    Department of Engineering, University of Liverpool,

    Received 5 October 2006; received in reviseAvailable onli

    Abstract

    Numerical calculations have been carried out for ow in a trunc(radius increasing with approach to the rotating endwall) and divangle of inclination of the sidewall. At the same time Moat eddthe sidewall and the non-rotating endwall. For the divergent geommerges with the Moat eddies. The ow phenomena identied frof pressure around the periphery of the computational domain.0142-727X/$ - see front matter 2007 Elsevier Inc. All rights reserved.doi:10.1016/j.ijheatuidow.2007.04.018tainer by a rotating endwall

    Leary, R.J. Poole

    wnlow Street, Liverpool L69 3GH, United Kingdom

    rm 16 April 2007; accepted 18 April 200713 June 2007

    d cone generated by rotation of one endwall. For both convergentent geometries, vortex breakdown is suppressed beyond a certainof increasing strength and extent appear in the corner betweeny, a zone of recirculation appears on the sidewall and eventuallystreamline patterns are consistent with the calculated variation

    www.elsevier.com/locate/ijh

    Flow 28 (2007) 14181428

  • Nomenclature

    a, b, p constants in accuracy estimation schemeCP coecient of pressureH height of conical container (m)p static pressure (Pa)pc stagnation pressure at centre of non-rotating

    endwall (Pa)pref reference pressure (static pressure at centre of

    rotating endwall) (Pa)r radius of container at distance x (m)R radius of rotating endwall (m)REQ volume equivalent radius

    R2 RH tan a H2 tan2 a=3q

    (m)

    RM mean radius of truncated cone R (H tana)/2

    XS non-dimensional distance along axis to rststagnation point

    M.P. Escudier et al. / Int. J. Heat andeect on the appearance of a single recirculation bubble.In a more recent paper, Mullin et al. (2000) showed bothexperimentally and numerically that the recirculation bub-ble was strongly inuenced by the introduction of a taperedcentrebody, which could be stationary or rotating. Theyargued that the sign of the axial pressure gradient dependedupon whether the centrebody diameter increased ordecreased between the non-rotating and the rotatingendwalls and whether it was stationary or rotating, andthat these two inuences had a strong inuence on themagnitude of the breakdown region. A decreasing diame-ter led to an adverse pressure gradient when the innercylinder was rotating and enlargement of the breakdownwhereas an increasing diameter had the opposite eect.The eects were also reversed if the inner cylinder wasstationary.

    (m)The investigation reported here is numerical and con-cerned with the ows produced in truncated conical con-

    A

    B C

    D

    H = 2R

    A

    B C

    D

    R

    rotatingendwall

    Fig. 1. Schematic of ow domain (a) divergent geometry, a > 0 and (b)convergent geometry, a < 0.tainers by a rotating endwall (illustrated schematically inFig. 1). The three global parameters required to character-ise the problem are the Reynolds number Re = XR2/t, theaspect ratio H/R, and the slope angle of the inclined wall,a, taken as positive for the divergent geometry (radius rincreasing with axial distance x from the rotating disk)and negative for the convergent geometry (r decreasing).Similarities are apparent between our geometries and theconical arrangement investigated by Buhler (1994) and alsowith the tapered centrebody geometries of Mullin et al.(2000) in the sense that the divergent geometry, a > 0, cor-responds with Mullin et al.s decreasing gap, while the con-vergent geometry, a < 0, corresponds with their increasinggap. An essential dierence compared with the geometriesinvestigated by Mullin et al., which precludes a direct com-parison between our results and theirs, is that in the

    a angle of inclination of container sidewallDr radial extent of computation cell (m)Dx axial extent of computation cell (m)g dynamic viscosity of uid (Pa s)t kinematic viscosity of uid (m2/s)q density of uid (kg/m3)X angular velocity of rotating endwall (s1)Re Reynolds number XR2/tReEQ Reynolds number XR2EQ=tReM Reynolds number XR2M=tu ow velocity (m/s)x distance along container axis from rotating disk

    (m)

    Fluid Flow 28 (2007) 14181428 1419absence of a centrebody the no-slip condition is imposedonly on the outer conical surface and the rotating endwall.As they showed, their results were critically dependent onwhether the inner cylinder was rotating not simply uponarea change. In addition to streamline patterns, we alsopresent the pressure variation over the surface of the con-tainer to further explore the pressure-gradient argumentsof Mullin et al.

    2. Numerical method

    To compute the ow eld within the rotating endwallconical container, discussed above, we make use of theassumption that the ow is laminar, incompressible, steadyand axisymmetric (i.e. two-dimensional). The governingequations are then those expressing conservation of massEq. (1) and momentum Eq. (2):

    r u 0; 1r quu rp gr ru: 2

  • In these equations, u represents the vector velocity, p thestatic pressure, q the uid density and g its dynamic viscos-ity. We used the commercial package FLUENT (version6.0) to solve Eqs. (1) and (2). This well-established codehas been used extensively in the calculation of complexows (see Fellouah et al., 2006; Huang et al., 2006; Tahaand Cui, 2006; Hu et al., 2005 for recent examples) and,with the correct implementation, is adequate to modelthe laminar ows under consideration here. We used thesegregated solver in which the momentum and swirl veloc-ity equations were discretised using a second-order up-winding scheme. Coupling of the pressure and velocitywas achieved using the well-known SIMPLEC implementa-tion of Van Doormal and Raithby (1984).

    Double precision (14 d.p.) was used for all the calcula-tions so that round-o errors are negligible. The iterationswere terminated whenever the scaled residuals (see Celikand Li, 2005) for the solutions for the two componentsof velocity and the continuity equation approached anasymptotic value. In general the scaled residuals wereobserved to reach a level between 1 1012 and 1 1015.

    A preliminary series of calculations was carried out with200 100, 400 200 and 800 400 uniform (i.e. Dr = Dx)

    cells for the base case (a = 0) to investigate the accuracyof our simulations. Firstly we made a qualitative compari-son between our calculated streamlines and the experimen-tal visualisations of Escudier (1984). Such a comparison isshown in Fig. 2 for the particularly complex case of a dou-ble breakdown. As can be seen the simulation predicts boththe occurrence of the primary and secondary recirculationbubbles as well as their size and location along the centre-line. In addition to this qualitative comparison, our consis-tent mesh-renement procedure allows us to undertake amore quantitative validation and to directly estimate thenumerical accuracy of our simulations. A sensitive criterionto determine this accuracy is the axial distance from thenon-rotating endwall to the location along the centrelineof the rst stagnation point, non-dimensionalised by thedisk radius, which we choose here to dene XS. The varia-tion of this quantity with increasing mesh renement isshown in Fig. 3. Firstly we note that the variation of XSbetween meshes is about 1% at most and so can beregarded as negligibly small. Secondly, tting these pointsto an equation of the form a(Dr)p + b allows us to estimatethe order of accuracy (p) of our simulations (Ferziger andPeric, 2001). Although our simulations are nominally sec-

    1420 M.P. Escudier et al. / Int. J. Heat and Fluid Flow 28 (2007) 14181428Fig. 2. Comparison of computed streamlines and experimental ow visualisatioRe = 1854, H/R = 2 (Mesh M2).n (taken from Escudier, 1984) of ow produced in a cylindrical container,

  • andr/R

    X S

    0 0.0025 0.005 0.0075 0.01 0.01250.2075

    0.2085

    0.2095

    0.2105

    0.2115

    0.2125

    -0.737(r/R)1.22 + 0.2112

    Fig. 3. Variation of non-dimensional distance to stagnation point XS withmesh renement for base case (a = 0).

    Table 1Mesh characteristics for base case together with estimates of numericalaccuracy

    Mesh (a = 0) NC XS % error in XSM1 20000 0.2085 1.31

    M.P. Escudier et al. / Int. J. Heatond order in accuracy, we note that they are in reality onlyslightly better than rst order (p = 1.2). However if we usethis order to estimate the Richardson extrapolation valuefor this quantity (i.e. the value extrapolated to zero meshsize), shown in Table 1, we still nd that the error in oursimulations is negligibly small.

    On the basis of this analysis of the mesh dependency, allsubsequent calculations were carried out with a nominal400 200 (i.e. 80000) cells. The actual number of cells(NC) used was determined by the geometry under consid-eration since the smaller computational domain of the con-vergent geometry (a < 0) required fewer cells whereas thedivergent geometry (a > 0) required more. In addition, aner mesh was used in the corner between the inclined side-wall and the non-rotating endwall to resolve the small-scaleow structure as discussed below. Sahin and Owens (2003),who investigated the two-dimensional, lid-driven cavityow discuss in some detail numerical problems arisingfrom the singularities which arise in the corners betweenthe sliding lid and the stationary cavity walls and the mea-sures implemented to overcome them. In our case, a similarsingularity occurs at the corner between the rotating diskand the container sidewalls but no special measures wereneeded to address numerical problems.

    The main parameter varied in the calculations was thecone angle a. Calculations were carried out for a between+25 and the limiting case of 26.6 while the aspect ratio

    M2 80000 0.2101 0.56M3 320000 0.2107 0.24Richardson extrapolation 0.2112 0.00was xed at H/R = 2. For the majority of calculations theReynolds number chosen was 1854 with a few additionalcalculations carried out for 2354 and 3354. The lower Rey-nolds number corresponds to a condition just within thebase-case (i.e. a = 0) boundary for which Escudier(1984) found a double breakdown whereas the higher Rey-nolds number is just above the boundary and correspondsto a single breakdown.

    3. Results of numerical calculations

    3.1. Streamline patterns for the convergent geometry

    As is immediately apparent from Fig. 4, for Re = 1854,breakdown is suppressed for a between 7 and 8although even for the limiting case of a = 26.6 a slightbulge is still evident in the streamlines in the vicinity ofthe breakdowns which occur for a > 8. The bulge movesprogressively along the axis, towards the rotating disk, asjaj increases. If we regard the geometry with a < 0 as cor-responding to gap width increasing towards the rotatingend, then this behaviour is qualitatively similar to whatwas found by Mullin et al. (2000) when the inner cylinderwas not rotating.

    A feature not remarked upon hitherto in investigationsof the container problem is the appearance of a Moat cor-ner-eddy system (Moat, 1964) in the corner between thesidewall and the non-rotating endwall. Such eddies havebeen observed in a number of other geometries involvingrotating surfaces including the concentric-cone arrange-ment investigated by Hall et al. (2007). This omission forthe container problem is surprising because the existenceof such vortices might have been anticipated. In fact, care-ful scrutiny of streamline patterns reported in several ear-lier papers (e.g. Lugt and Haussling, 1982; Bhattacharyaand Pal, 1998; Brons et al., 1999) hints at their presencethough it may be that the earlier calculations were per-formed with grids too coarse to conrm the occurrenceof Moat vortices. The analytical approach adopted byHall et al. (2007) reveals the rst four of an innitesequence of counter-rotating toroidal eddies as the apexof their cones is approached. As can be seen in Fig. 4,the scale and penetration of the Moat eddies increaseswith increasing values of jaj but it is clear that a Moateddy is present for all values of a, including 0.

    For the higher Reynolds number of 2354, shown inFig. 5, the streamlines are qualitatively similar to thosefor Re = 1854. Breakdown is suppressed for a < 12 com-pared with 8 for the lower Re value while the Moateddies appear more pronounced at about the same inclina-tion angle.

    3.2. Streamline patterns for the divergent geometry, a > 0

    The ow behaviour for a > 0 as revealed in Fig. 6 is

    Fluid Flow 28 (2007) 14181428 1421somewhat more complex than for the convergent cases.The breakdown is again gradually suppressed as a

  • and1422 M.P. Escudier et al. / Int. J. Heatincreases but for Re = 1854 is still clearly evident for thehighest inclination angle, 25, and for a greater than 10Moat eddies appear in the upper corner formed betweenthe inclined sidewall and the non-rotating endwall. Fora = 20 a second Moat eddy is just detectable and fora = 25 is very well resolved by the calculation. Anotherprominent feature of the calculations is the toroidal blis-ter of recirculation which appears high up on theinclined sidewall for a > 20. As we shall see from calcu-lations of the pressure distribution, this separation bubbleis associated with the occurrence of a positive pressuregradient on the inclined wall. One of the cases reportedby Hall et al. (2007) for their concentric-cone geometry

    Fig. 4. Streamlines for convergent geometry with Re = 1854, (a) a = 0, (bFluid Flow 28 (2007) 14181428also reveals a zone of recirculation on the outer cone wallfor the situation where the cones are contra rotatingwhile the spherical lid rotates in the same sense as theinner cylinder.

    From Fig. 7 it is seen that at the higher Reynolds num-ber of 2354 breakdown on the axis is completely suppressedfor a > 10 although the near-axis streamlines remainhighly distorted even for the highest inclination angle.Moat eddies are easily seen even for a = 5, the onset ofsidewall separation is just apparent for a = 20, and by25 has grown considerably and is on the verge of mergingwith the Moat eddies. For the nal calculation, the resultsof which are shown in Fig. 8, with a = 25 and Re = 3354,

    ) a = 7, (c) a = 8, (d) a = 23, (e) a = 24 and (f) a = 26.6.

  • andM.P. Escudier et al. / Int. J. Heatthe Moat eddy system has indeed merged with theinclined-wall recirculation bubble. For this Reynolds num-ber, as shown by the experiments of Escudier (1984), thereis no breakdown even for the base case, a = 0 and it islikely that in reality the ow would be unsteady.

    The calculations presented in this paper are all for thesame value of an aspect ratio dened as H/R. It might bethought that a more appropriate measure of the aspectratio would be one dened in terms of either the meanradius RM = R (H tana)/2 or a volume-equivalent radiusREQ

    R2 RH tan a H 2 tan2 a=3

    q. It has been sug-

    gested that the boundaries for the occurrence of vortex

    Fig. 5. Streamlines for convergent geometry with Re = 2354, (a) a = 0, (b)Fluid Flow 28 (2007) 14181428 1423breakdown in terms, of H/RM or H/REQ versus Reynoldsnumber (dened as either Re = XR2/t, ReM XR2M=t orReEQ XR2EQ=t) might correspond to the boundaries forow in a cylindrical container, as identied experimentallyby Escudier (1984). To explore this possibility, in Fig. 9 wehave plotted points corresponding to the numerical calcu-lations presented in this paper using the volume-equivalentradius REQ in the denitions of both the aspect ratio andthe Reynolds number (thin curves). Also shown (alongthe horizontal broken lines) are points in which H/R hasbeen modied but not Re. A similar plot based on RMfor the characteristic radius led to very small dierencesand is not shown here.

    a = 12, (c) a = 13, (d) a = 23, (e) a = 24 and (f) a = 26.6.

  • ab

    c

    Fig. 6. Streamlines for divergent geometry with Re = 1854, (a) a = 5, (b) a = 20 and (c) a = 25.

    1424 M.P. Escudier et al. / Int. J. Heat and Fluid Flow 28 (2007) 14181428

  • anda

    M.P. Escudier et al. / Int. J. HeatFor Re = 1854, the closed points Fig. 9 correspond toa = 25, 20, 5, 0 and 7, while the open points corre-spond to 8, 23, 24 and 26.6. For Re = 2354,the open points to the left correspond to a = 25, 20 and15, those to the right correspond to 13, 23, 24

    c

    e

    Fig. 7. Streamlines for divergent geometry with Re = 2354, (a)b

    Fluid Flow 28 (2007) 14181428 1425and 26.6 while the closed points correspond to 10, 5and 12. It is clear that the proposed criterion fails forabout one third of the cases considered here (some pointslie outside the range of the graph), both in terms ofpredicting breakdown where it does not occur and not

    d

    a = 5, (b) a = 10, (c) a = 15, (d) a = 20 and (e) a = 25.

  • and1426 M.P. Escudier et al. / Int. J. Heatpredicting it where it does. More specically, at the lowerReynolds number the criterion predicts breakdown fora = 25 where none is observed but not for 8 where itis (in fact the latter case falls on the one/two breakdownboundary). For the higher Reynolds number, the criterionworks well for a > 0 but fails badly for a = 13 where nobreakdown is predicted while in practice this is well withinthe boundary for two breakdowns to occur.

    The motivation for dening an eective radius was thepossibility that the eect of sidewall inclination could beexplained in terms of a modied aspect ratio. Althoughthe explanation is not fully supported by our calculations,

    calculations if only H/R is modied.

    Fig. 8. Streamlines for divergent geometry wi

    H/R

    Re

    1 1.5 2 2.5 3 3.51000

    1500

    2000

    2500

    3000

    3500

    4000

    2 BREAKDOWNS

    1 BREAKDOWN

    Fig. 9. Closed symbols, j and m, correspond to vortex breakdownaccording to the numerical calculations; open symbols, h and n, to theabsence of breakdown. The thick curves correspond to the boundariesestablished experimentally by Escudier (1984).3.3. Wall pressure distribution

    Mullin et al. (2000) demonstrated that for their geome-try enlargement or suppression of vortex breakdown onit could be argued that the general trend indicated by thepoints in Fig. 9 suggests that part of the inuence of theinclined sidewalls is an eective change in aspect ratio.The breakdown criterion is in better agreement with the

    th Re = 3354, (a) a = 0 and (b) a = 25.

    Fluid Flow 28 (2007) 14181428the surface of the central conical rod which they investi-gated was associated with a pressure gradient broughtabout by convergence or divergence of the annular gapand rotation or non-rotation of the rod. When the innercylinder was rotating and tapered towards the rotating cyl-inder, an adverse pressure gradient was generated whichenhanced the breakdown region whereas breakdown wasdiminished when the inner cylinder increased in diameter.The trends were reversed in the absence of rotation. Aswe have already commented, although there is clearly anassociation between area change and changes in break-down structure in our calculations, there are essential dif-ferences between our work and that of Mullin et al. as aconsequence of the absence of a centrebody to provide ano-slip boundary condition for our problem so direct com-parisons cannot be made.

    The variation of the pressure coecient, CP 2(p pref)/qX2R2, with distance s around the periphery ofa central plane is shown in Figs. 1012. The reference pres-sure pref has been taken as the pressure at the centre of therotating disk (i.e. location B in Fig. 1). Fig. 10 shows thevariation of CP from the centre of the non-rotating endwall(A), down the axis to the rotating disk (B), and part wayacross the rotating disk. The curves are shown for 5 inter-

  • s/R

    C P

    0 0.25 0.5 0.75 1 1.25 1.5 1.75 2 2.25-0.012

    -0.008

    -0.004

    0

    0.004

    BA 0.2 R

    25

    A

    0-5

    -10

    -15

    limit

    -20-25

    Fig. 10. Variation of pressure coecient around domain boundary for0 < s < 2.25R. The curves correspond to 5 intervals in the inclinationangle a plus the limiting case.

    s / R

    C P

    2 2.25 2.5 2.75 3 3.25 3.5

    0

    0.05

    0.1

    0.15

    0.2CB

    limit

    25

    Fig. 11. Variation of pressure coecient across rotating endwall and partway along inclined sidewall.

    C P

    3.25 3.5 3.75 4 4.25 4.5 4.75 5 5.25 5.5 5.75 6

    0

    0.02

    0.04

    0.06A

    limit

    D

    025

    C P

    3.25 3.5 3.75 4 4.25 4.5 4.75 50.015

    0.02

    0.025

    0.03

    0.035

    Fig. 12. Variation of pressure coecient along inclined sidewall andacross non-rotating endwall. The non-dimensional distance from C to Dhas been scaled to range from 3 to 5 and that from D to A to range from 5to 6.

    M.P. Escudier et al. / Int. J. Heat andvals in a from +25 to 25 plus the limiting case (26.6).In this gure the distance s has been normalised with theradius R. Although it is the pumping action of the rotatingdisk which produces ow within the container, in fact thepressure dierence (pc pref), pc being the stagnation pres-sure at the centre of the non-rotating disk (A), is negativefor 15 > a > 5 and only becomes positive for a > 15and a < 5. For the divergent geometry the pressure fallsrapidly for 0 < s < 0.2R before increasing to an internalstagnation point (corresponding to the nose of thebreakdown region) at about 0.4R where the CP curvesreveal a pronounced knee. The changes for the conver-gent case are more complex and show a considerable pres-sure increase at A for a < 5 combined with a smootherrecovery beyond the minimum. The absence of a welldened knee in the curve corresponds to the suppressionof vortex breakdown for a < 5. For 1.4 < s/R < 2, withapproach to the spinning disk, the CP curves for all inclina-tion angles are practically indistinguishable.

    The variation of CP across the rotating endwall (B to C)and part way along the inclined sidewall is shown in Fig. 11with s again scaled with R. The pressure increases to a peakat the periphery of the spinning disk, much as would beexpected though the pressure levels for all cases are consid-erably below that for solid-body rotation (i.e. CP = (r/R)

    2).The monotonic changes in the peak and the curves gener-ally suggest that as a decreases, the sidewalls increasinglyact essentially as a blockage.

    In the nal gure showing the variation of CP (Fig. 12),the distance along the domain boundary has been scaled tohave the value 3 at location C (the edge of the rotating end-wall), the value 5 and at location D and the value 6 at loca-tion A. Thus distance along CD has been scaled with itslength 2R seca and along DA with its length R + 2Rco-seca. The eect of this normalisation is that the computa-tional domain has been eectively transformed to arectangle. The most signicant feature of Fig. 12 is thechange in pressure gradient from negative to positive (orfavourable to adverse) on the inclined sidewall for abetween 10 and 15. For a = 25 the gradient is sucientlyadverse to cause the separation blister seen in Fig. 4. Thetwo lled circles on the 25 curve in the larger scale insertin Fig. 12 delineate the extent of the blister.

    4. Conclusions

    The results of numerical calculations have been pre-sented for the ow generated in a truncated cone by rota-tion of one endwall. For an aspect ratio H/R = 2 and aReynolds number Re of 1854, the results show that vortexbreakdown which occurs for the base case of a circular cyl-inder is suppressed if the angle of inclination of the conicalsidewall a is less than 7, for the convergent geometry,but is still present, although reduced in magnitude, fora = 25 for the divergent geometry. The results are qualita-

    Fluid Flow 28 (2007) 14181428 1427tively similar to those reported by Mullin et al. (2000) for astationary inner cylinder. Moat eddies are increasingly

  • prominent in the corner between the sidewall and the non-rotating endwall as the inclination angle is increased. Forthe divergent geometry with 25 inclination angle, a furtherregion of recirculation appears about three quarters of theway along the sidewall as a result of the adverse pressuregradient which arises for a > 10. Qualitatively similarresults are found at higher Reynolds numbers althoughthe breakdown is also suppressed in the divergent geometryif a > 10 for Re = 2354, and for the combinationRe = 3354 with a = 25 the Moat eddies merge with therecirculation region. For some conditions, the calculationsof Hall et al. (2007) reveal a similar recirculation blister onthe outer wall of their concentric-cone geometry.

    Acknowledgements

    Fellouah, H., Castelain, C., Ould El Moctar, A., Peerhossaini, H., 2006. Anumerical study of Dean instability in non-Newtonian uids. J. FluidsEng. Trans. ASME 128, 3441.

    Ferziger, J.H., Peric, M., 2001. Computational Methods for FluidDynamics. Springer, New York.

    Fujimura, K., Koyama, H.S., Hyun, J.M., 2004. An experimental studyon vortex breakdown in a dierentially-rotating cylindrical container.Exp. Fluids 36, 399407.

    Hall, O., Hills, C.P., Gilbert, A.D., 2007. Slow ow between concentriccones. Q. J. Mech. Appl. Math. 60 (1), 2748.

    Hu, L.Y., Zhou, L.X., Zhang, J., Shi, M.X., 2005. Studies on stronglyswirling ows in the full space of a volute cyclone separator. AICHE J.51 (3), 740749.

    Huang, Z., Olsen, J.A., Kerekes, R.J., Green, S.I., 2006. Numericalsimulation of the ow around rows of cylinders. Comp. Fluids 35, 485491.

    Lopez, J.M., 1990. Axisymmetric vortex breakdown: Part 1. Connedswirling ow. J. Fluid Mech. 221, 533552.

    1428 M.P. Escudier et al. / Int. J. Heat and Fluid Flow 28 (2007) 14181428The authors thank Professor Manuel Alves, Faculdadede Engenharia Universidade do Porto, for advice regardingthe estimation of the numerical accuracy of our simula-tions. We are also grateful to a referee for suggesting theuse of a volume-equivalent or mean radius to explore theeect of aspect ratio on the occurrence of breakdown andto another referee for bringing two relevant references toour attention.

    References

    Bhattacharya, S., Pal, A., 1998. Axisymmetric vortex breakdown in a lledcylinder. Int. J. Eng. Sci. 36 (5/6), 555563.

    Brons, M., Voigt, L.K., Sorensen, J.K., 1999. Streamline topology ofsteady axisymmetric vortex breakdown in a cylinder with co- andcounter-rotating end-covers. J. Fluid Mech. 401, 275292.

    Buhler, K., 1985. Stromungsmechanische Instabilitaeten zaher Medien imKugelspalt, Fortschrittberichte VDI Reihe 7(96).

    Buhler, K., 1994. Wirbelstromungen mit Ruckstromgebiten. DGLRJahrbuch II, 887894.

    Celik, I.B., Li, J., 2005. Assessment of numerical uncertainty for thecalculations of turbulent ow over a backward-facing step. Int. J.Numer. Meth. Fluids 49 (9), 10151031.

    Escudier, M.P., 1984. Observations of the ow produced in a cylindricalcontainer by a rotating endwall. Exp. Fluids 2, 189196.Lugt, H.J., Haussling, H.J., 1982. Axisymmetric vortex breakdown inrotating uid within a container. Trans. ASME J. Appl. Mech. 49,921923.

    Moat, H.K., 1964. Viscous and resistive eddies near a sharp corner. J.Fluid Mech. 18, 118.

    Mullin, T., Tavener, S.J., Clie, K.A., 1998. On the creation of stagnationpoints in a rotating ow. J. Fluids Eng. 120, 685689.

    Mullin, T., Kobine, J.J., Tavener, S.J., Clie, K.A., 2000. On the creationof stagnation points near straight and sloped walls. Phys. Fluids 12 (2),425431.

    Pereira, J.C.F., Sousa, J.M.M., 1999. Conned vortex breakdowngenerated by a rotating cone. J. Fluid Mech. 385, 287323.

    Sahin, M., Owens, R.G., 2003. A novel fully implicit nite volume methodapplied to the lid-driven cavity problem Part I: High Reynoldsnumber ow calculations. Int. J. Numer. Meth. Fluids 42, 5777.

    Spohn, A., Mory, M., Hopnger, E.J., 1998. Experiments on vortexbreakdown in a conned ow generated by a rotating disc. J. FluidMech. 370, 7399.

    Taha, T., Cui, Z.F., 2006. CFD modeling of slug ow in vertical tubes.Chem. Eng. Sci. 61, 676687.

    Tsitverblit, N., Kit, E., 1996. Numerical study of axisymmetric vortexbreakdown. Acta Mech. 118, 7998.

    Van Doormal, J.P., Raithby, G.D., 1984. Enhancements of the SIMPLEmethod for predicting incompressible uid ows. Numer. HeatTransfer 7, 147163.

    Vogel, H.U., 1968. Experimentelle Ergebnisse uber die laminare Stromungin einem zyindrischen Gehause mit darin rotierende Scheibe, MPI furStromingsforschung, Bericht 6.

    Flow produced in a conical container by a rotating endwallIntroductionNumerical methodResults of numerical calculationsStreamline patterns for the convergent geometryStreamline patterns for the divergent geometry, alpha gt 0 deg Wall pressure distribution

    ConclusionsAcknowledgementsReferences