+ All Categories
Home > Documents > Flow along a diverging channel

Flow along a diverging channel

Date post: 11-Oct-2016
Category:
Upload: m-b
View: 221 times
Download: 3 times
Share this document with a friend
21
Journal of Fluid Mechanics http://journals.cambridge.org/FLM Additional services for Journal of Fluid Mechanics: Email alerts: Click here Subscriptions: Click here Commercial reprints: Click here Terms of use : Click here Flow along a diverging channel S. C. R. DENNIS, W. H. H. BANKS, P. G. DRAZIN and M. B. ZATURSKA Journal of Fluid Mechanics / Volume 336 / April 1997, pp 183 202 DOI: 10.1017/S0022112096004648, Published online: 08 September 2000 Link to this article: http://journals.cambridge.org/abstract_S0022112096004648 How to cite this article: S. C. R. DENNIS, W. H. H. BANKS, P. G. DRAZIN and M. B. ZATURSKA(1997). Flow along a diverging channel. Journal of Fluid Mechanics, 336, pp 183202 doi:10.1017/S0022112096004648 Request Permissions : Click here Downloaded from http://journals.cambridge.org/FLM, IP address: 139.184.30.131 on 24 Sep 2012
Transcript
  • JournalofFluidMechanicshttp://journals.cambridge.org/FLM

    AdditionalservicesforJournalofFluidMechanics:

    Emailalerts:ClickhereSubscriptions:ClickhereCommercialreprints:ClickhereTermsofuse:Clickhere

    Flowalongadivergingchannel

    S.C.R.DENNIS,W.H.H.BANKS,P.G.DRAZINandM.B.ZATURSKA

    JournalofFluidMechanics/Volume336/April1997,pp183202DOI:10.1017/S0022112096004648,Publishedonline:08September2000

    Linktothisarticle:http://journals.cambridge.org/abstract_S0022112096004648

    Howtocitethisarticle:S.C.R.DENNIS,W.H.H.BANKS,P.G.DRAZINandM.B.ZATURSKA(1997).Flowalongadivergingchannel.JournalofFluidMechanics,336,pp183202doi:10.1017/S0022112096004648

    RequestPermissions:Clickhere

    Downloadedfromhttp://journals.cambridge.org/FLM,IPaddress:139.184.30.131on24Sep2012

  • J. Fluid Mech. (1997), vol. 336, pp. 183{202

    Copyright c 1997 Cambridge University Press183

    Flow along a diverging channel

    By S. C. R. D E N N I S1, W. H. H. B A N K S2,P. G. D R A Z I N2 AND M. B. Z A T U R S K A2

    1Department of Applied Mathematics, University of Western Ontario, London, Ontario,Canada N6A 5B9

    e-mail: [email protected]

    2School of Mathematics, University of Bristol, Bristol BS8 1TW, UKe-mail: [email protected]; [email protected]; [email protected]

    (Received 27 February 1995 and in revised form 29 October 1996)

    This paper treats the two-dimensional steady flow of a viscous incompressible fluiddriven through a channel bounded by two walls which are the radii of a sector and twoarcs (the inlet and outlet), with the same centre as the sector, at which inflow andoutflow conditions are imposed. The computed flows are related to both a laboratoryexperiment and recent calculations of the linearized spatial modes of Jeery{Hamelflows. The computations, at a few values of the angle between the walls of the sectorand several values of the Reynolds number, show how the rst bifurcation of the flowin a channel is related to spatial instability. They also show how the end eects due toconditions at the inlet and outlet of the channel are related to the spatial modes: inparticular, Saint-Venants principle breaks down when the flow is spatially unstable,there being a temporally stable steady flow for which small changes at the inlet oroutlet create substantial eects all along the channel. The choice of a sector as theshape of the channel is to permit the exploitation of knowledge of the spatial modesof Jeery{Hamel flows, although we regard the sector as an example of channels withwalls of moderate curvature.

    1. IntroductionThe flow of a viscous incompressible fluid in a two-dimensional channel poses

    a classical problem discussed in many textbooks (e.g. Batchelor 1967, x5.6). Theproblem has a wide range of engineering and environmental applications. It has beenstudied by use of mathematical, numerical and experimental methods, yet even theearly stages of the bifurcations on the route to turbulence are poorly understood.

    Asymptotic methods may be used to elucidate the bifurcations and instabilities ofthe channel flows. One asymptotic approach (cf. Georgiou & Eagles 1985 and earlierpapers by Eagles and co-authors) is to assume that the walls of the channel arenearly parallel so that the flow is locally parallel and its instability may be found tothe rst approximation by solving an Orr{Sommerfeld problem. A similar approach,due to Fraenkel (1962), is to assume that the curvature of the walls of the channelis small so that the flow is locally radial, i.e. a Jeery{Hamel flow. (Recall thata Jeery{Hamel flow is a steady two-dimensional radial flow between two inclinedrigid planes driven by a line source at the intersection of the planes, and is describedexactly by a similarity solution of the Navier{Stokes equations.) We shall mentionJeery{Hamel flows so often in the paper that it will be useful to call them JH flows.

  • 184 S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska

    At this stage it is convenient to dene the Reynolds number R in a form suitablefor use throughout the paper, namely R = Q=2, where Q is the volume flux per unitdistance normal to the plane of flow, and is the kinematic viscosity of the fluid.Thus we suppose that the flow is driven steadily with flux Q > 0. The advantage ofthis denition is that it does not directly involve a length scale of the channel, and sois suitable for all two-dimensional channel flows driven steadily. There is nothing veryspecial about the choice of the factor 1

    2, but it will simplify some of our arithmetic.

    On using the approximation of nearly parallel flow, it is found that instability arisesat a large value of R. The mode of instability is a travelling wave, i.e. the principleof exchange of stabilities is invalid; so it may be presumed that there is a Hopfbifurcation with the onset of a time-periodic flow as the Reynolds number increases(although the time-periodic flow may be subcritical and so unstable).

    Banks, Drazin & Zaturska (1988) followed Fraenkels approach and went on toconsider the spatial and temporal stability of Jeery{Hamel flows. The analysis byBanks et al. of steady spatial perturbations of a JH flow, developing the pioneeringideas of Dean (1934), suggests that the flow is spatially stable for values of theReynolds number R < R2(), where is the semi-angle between the plane walls andR2 is a certain known function. It is similarly stable to symmetric perturbations whenR < R3(). In practice the flow may be simulated in the laboratory by flow in achannel shaped like a sector with an inlet at an arc of small radius and an outlet atan arc of large radius. Then the analysis suggests that when R > R2() the inlet andoutlet conditions aect the whole steady flow, i.e. even if the inlet were at a very smallradius and the outlet at a very large radius, there would be signicant end eectsthroughout the domain of flow, and, further, no JH flow would be observed locallyat any station. The value R2() of the Reynolds number at the onset of instability is amoderate one (unless is small), and the mode of instability is a small rotation of thebasic flow, the principle of exchange of stabilities being valid; there is a correspondingsubcritical pitchfork bifurcation, such that there is an unstable symmetric solution forsmall R R2() > 0 and a stable symmetric solution and two unstable asymmetricsolutions for small R R2() < 0.

    The words spatial stability are used so often in this paper that it is appropriateto note that they concern not so much stability, which is usually meant to describethe evolution of perturbations in time, as the behaviour in space of steady flowsnear to the basic steady flow under consideration. The close relationship of thebifurcation of steady flows to their temporal instability justies the common usageof the phrase spatial stability, even when it strictly concerns only the developmentof steady flows in space. So spatial stability determines stability indirectly. However,it should not be forgotten that our concentration on steady flows and their spatialstability gives no information about temporal instabilities at Hopf bifurcations, andthat such instabilities may be important physically, especially at small values of for which R2 is large, and it is known that parallel flows have a subcritical Hopfbifurcation (cf. Drazin & Reid 1981, Chap. 4).

    If a channel has walls which are both nearly parallel and of small curvature then wepresume that both the Orr{Sommerfeld and Jeery{Hamel modes of instability mayoccur. The one which would occur in practice as the Reynolds number is increasedvery slowly seems to depend upon the details of the conguration of the channel,such as the small angle between the walls and the small curvature.

    Clie & Greeneld (1982), Sobey & Drazin (1986) and Fearn, Mullin & Clie (1990)have found the steady flows in various channels by direct numerical integration of thegoverning Navier{Stokes equations; Sobey & Drazin also found some time-periodic

  • Flow along a diverging channel 185

    Figure 1. Sketch based on Nakayamas experimental results showing hydrogen bubble proles:R = 300; = =18.

    flows. In each of the three papers it is found that for a symmetric channel thereis a unique stable flow for small values of the Reynolds number R, but that asR increases above a certain critical value, Rc say, the flow becomes unstable andthere is a supercritical pitchfork bifurcation such that for small R Rc > 0 thereare two stable asymmetric steady flows as well as the unstable symmetric steadyflow. The symmetric channels considered in these papers are rather dierent, yet it issurprising that Sobey & Drazin found Rc to be much smaller than the other authorsfound it to be; Sobey & Mullin (1992) suggested that Sobey & Drazins values areinaccurate because of their use of upwind dierences. Another surprising result isthat the pitchfork bifurcation found in these papers on computational fluid dynamicsis supercritical yet the pitchfork bifurcation of JH flows which is associated with theirinstability and symmetry breaking is subcritical at R = R2 (recall that the JH flow isdescribed by a similarity solution); it follows that there is no JH flow of the computedtype to approximate the flow locally in a channel with walls of small curvature whenR > Rc. So it appears that JH flows are irrelevant to flow in a diverging channelwith walls of small curvature if R > Rc or, at least, irrelevant where there is no stablelocal JH flow. The exact symmetry of the channel is not crucial to these conclusions,although the pitchfork bifurcation is an idealization of an exactly symmetric channel(cf. Sobey & Drazin 1986; Fearn et al. 1990). Borgas & Pedley (1990) have also usedsimilarity solutions to nd supercritical bifurcations in an indented channel.

    Hamadiche, Scott & Jeandel (1994) have recently investigated the temporal stabilityof some of the symmetric JH flows. They posed the linearized eigenvalue problem ona nite sector and proceeded to determine the critical value of the Reynolds numbercorresponding to marginal stability for various values of the semi-angle . Theyfound that loss of stability is supercritical (when is not very small) although theirmarginal curves are in surprisingly good agreement with the solutions of Sobey &Drazin (1986, x3) of the model stability problem of Hooper, Duy & Moatt (1982).

    There have been many laboratory experiments on flows in diverging channels atvarious values of the Reynolds number R, but few (e.g. Sobey & Drazin 1986; Fearnet al. 1990) have been carefully related to theoretical results. Nakayama (1988, gure105) has published a single photograph of flow along a diverging channel with planewalls at a semi-angle of = 10o, i.e. = =18 radians, at R = 300. We have prepareda sketch of Nakayamas experimental results from his gure 105 and present them ingure 1. This is a model of a diuser, as if designed to represent a JH flow: yet theflow shown has an asymmetric velocity prole with reverse flow near each wall, andno such JH flow exists. Again, we are drawn to the conclusion that, even if of all thelocal JH flows which exist none is stable, the flow in a channel may be steady andstable yet of a form dierent from a JH solution.

  • 186 S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska

    The aim of this paper is to clarify some of these diculties by use of computationalfluid dynamics. By judicious choice of problems and careful numerical methods wehave tried to bridge the gap between the mathematical results and previous numericaland experimental results, and resolve some of the paradoxes we have just described.In particular, we conrm the conjecture of Banks et al. (1988) that of the two familiesof eigenvalues known analytically for R = 0, one is associated with the growth ordecay for increasing r and the other for decreasing r.

    2. Flow in a sector2.1. Formulation of problem

    On the basis of the preceding discussion, we recognize that the problem of two-dimensional flow in a channel, whose shape is a sector with an inlet at an arc andan outlet at another arc, is fundamental to understanding flows in diverging channelswith small curvature. Banks et al. (1988) posed this problem mathematically, butBuhler & Kruckels (1990) initiated its numerical solution; however, their calculationswere few, made in short channels, and made for flows for which an unstable JHvelocity is imposed at the outlet although a stable JH flow exists at the values of R; used. Accordingly, we shall now formulate the problem, as illustrated in gure 2, andproceed to solve it numerically. Take plane polar coordinates (r; ) and correspondingvelocity components of the fluid as u; v respectively. We suppose that a steady flow isdriven by a given volume flux Q between the rigid impermeable walls with equations = , and that the inlet arc has equation r = r1 and the outlet arc r = r2 > r1. It isconvenient to use a streamfunction such that u = @ =r@ and v = @ =@r. Thenthe vorticity may be written as

    = r2 (2.1)and the Navier{Stokes equations reduce to the vorticity equation,

    1

    r

    @(; )

    @(r; )= r2; (2.2)

    where the Laplacian in cylindrical polar coordinates is r2 = @2=@r2 +@=r@r+@[email protected] the boundary conditions are

    = 12Q;

    @

    @= 0 at = (2.3)

    respectively on the sidewalls,

    = f1();@

    @r= g1() at r = r1 (2.4)

    at the inlet, and

    = f2();@

    @r= g2() at r = r2 (2.5)

    at the outlet, where the choice of f1(); g1(); f2(); g2() (such that f1() = f2() = 1

    2Q; f01() = f02() = g1() = g2() = 0) species the inlet and outlet velocities

    consistently with the given flux. We also consider in x3 replacing the condition on@ =@r at r = r1; r2 by a condition on , and in x4 discuss the relationship of the inletand outlet conditions to laboratory experiments.

    It is, of course, convenient to use dimensionless variables. So write r = r1r0; r2 =

    r1r02; =

    12Q 0; = Q 0=2r21 ; and rescale the inlet and outlet conditions. Then drop

  • Flow along a diverging channel 187

    O

    r

    v

    u

    r=r1

    h

    r=r2

    h=

    h=

    Figure 2. Sketch of the conguration of flow in a sector.

    the primes so that (2.1) remains unchanged and the problem reduces to

    1

    r

    @(; )

    @(r; )= R1r2; (2.6)

    = 1; @ @

    = 0 at = ; (2.7)

    = f1();@

    @r= g1() at r = 1; = f2();

    @

    @r= g2() at r = r2; (2.8)

    where the Reynolds number is dened as

    R = Q=2: (2.9)

    2.2. Jeery{Hamel flows

    Before describing our numerical method for solving (2.1), (2.6), (2.7) and (2.8), werecall briefly the results of the Jeery{Hamel similarity solution. For this, boundaryconditions (2.8) are ignored and a steady solution is sought in the form

    = G(y; ; R); (2.10)

    where y = =. Then problem (2.1), (2.6), (2.7) reduces to

    Giv + 42G00 + 2RG0G00 = 0; (2.11)

    G = 1; G0 = 0 at y = 1; (2.12)where a prime denotes dierentiation with respect to y.

    Much is known about these exact solutions, G(y; ; R), of the Navier{Stokes equa-tions and a summary of the results is given by the diagram in gure 3: fuller detailsare given by Fraenkel (1962), but they are too complicated to state here apart from

  • 188 S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska

    "3

    "2

    "1 "5

    "4

    "2

    "1

    1

    2

    3

    2.25

    =pi

    R0 5 105

    (a)

    (b)

    O OI or II1 II2"3

    Figure 3. (a) Boundaries Bn for some types of Jeery{Hamel solutions in the (R; )-plane (afterBanks et al. 1988, gure 3). Solutions of type I arise for values of (R; ) lying between the R-axisand curve B1, of type IIn between the curves Bn1 and Bn, and symmetric solutions of type IIInwith net inflow arise for R < 0. (b) Sketches of velocity proles of flows of types I; II1, and II2.

    the chief points relevant to the present work. It will be convenient to note that thepressure can be written as

    p =2(G0 + C)Rr2

    ; (2.13)

    where C is a constant, on absorbing any additional constant in the initial non-dimensionalization. There is an innity of solutions, both symmetric and antisymmet-ric, for any given pair of values of R; . Fraenkel denoted the types of the solutionsby the letters I; IIn; IIIn; IVn; Vn, where n is a counter for the number of maxima andminima of the velocity proles belonging to each type. The regions of the (R; )-planewhere a few of the solutions occur is indicated in gure 3(a), and the velocity prolesof the most relevant solutions I; II1; II2 are sketched in gure 3(b).

    Various perturbations of the JH solutions have been studied, but here we shallstudy the spatial development of such steady flows and their steady perturbationsafter Banks et al. (1988) and Goldshtik, Hussain & Shtern (1991); on writing

    (r; ; ; R) = G(y; ; R) + r ^(y; ; R); (2.14)

    substituting into (2.1), (2.6) and (2.7), and linearizing, an eigenproblem is derivedwith value = (; R). Little is known of the properties of (; R) for R 6= 0 but(; 0) is known analytically: there are two families. Each family can be divided intotwo innite subfamilies, with one subfamily of antisymmetric modes and another ofsymmetric modes, say a+;

    s+ respectively with Re() > 1 and

    a; s with Re() < 1.We have conjectured that the subfamilies a+;

    s+ represent the spatial growth (or

    decay) of steady disturbances as r decreases, while the subfamilies a; s representthe growth of disturbances as r increases. We also conjecture that the same subfamiliesoccur when R 6= 0 although their quantitative details dier from those when R = 0.

    Finally we note that the leading or most dangerous mode of the subfamiliesa+;

    s+, dangerous only in the sense that it grows most rapidly as r decreases, is

    antisymmetric, i.e. of all the modes belonging to a+; s+ the one with the minimum

  • Flow along a diverging channel 189

    value of Re() belongs to a+. Similarly, the mode of the subfamilies a; s with

    maximum value of Re() is antisymmetric. So we anticipate that flow in a half-sector0 6 y 6 1 (with symmetry boundary conditions on the symmetry plane y = 0) maybe spatially stable whereas flow in a full sector may be unstable for the same valueof R.

    3. Numerical method for steady flowNext we shall briefly describe the numerical formulation which we used to solve

    equations (2.1), (2.6), reverting to dimensional form.The sectorial region r1 6 r 6 r2; 6 6 of gure 2 is divided into the polar

    grid

    r = r1 + jh1; for j = 0; 1; : : : ; J; = + kh2 for k = 0; 1; : : : ; K; (3.1)where h1 = (r2 r1)=J; h2 = 2=K . At any node (r; ) of this grid, the spatialderivatives which appear in (2.1) and (2.6) are approximated by the usual central-dierence formulae. This gives the respective approximations to (2.1), (2.6) at thispoint as

    (1 + h1=2r) (r + h1; ) + (1 h1=2r) (r h1; ) + (2=r2)[ (r; + h2) + (r; h2)]2(1 + 2=r2) (r; ) + h21(r; ) = 0; (3.2)

    [1+h1=2r (R=4r)f (r; + h2) (r; h2)g](r + h1; )+[1 h1=2r + (R=4r)f (r; + h2) (r; h2)g]((r h1; )+[2=r2 + (R=4r)f (r + h1; ) (r h1; )g](r; + h2)+[2=r2 (R=4r)f (r + h1; ) (r h1; )g](r; h2)2(1 + 2=r2)(r; ) = 0; (3.3)

    where = h1=h2.Equations (3.2), (3.3) form two sets of equations which are obtained by setting

    (r; ) equal to the coordinates of each grid point in the region, and these mustbe solved numerically. The given boundary conditions for are of Dirichlet typefor (3.2), but it is necessary to calculate boundary values for by using all theconditions in (2.7), (2.8). Both the approximations (3.2), (3.3) are h2-accurate, so itis desirable to obtain h2-accurate approximations to the boundary values of . Wehave done this by following the procedure of Woods (1954) in which is expandedas a Taylor series in the appropriate variable normal to the boundary concerned.The corresponding normal gradient condition is used in this expansion, and higherderivatives are determined from equation (2.1); this leads to an approximation to theboundary value of in terms of internal values of , at the nearest grid point alongthe appropriate normal direction.

    The procedure is illustrated for the case of the boundary conditions at r = r1; r2. Ifwe take r = r1 as typical, we obtain

    (r1 + h1; ) = f1() + h1g1() +12h21

    @2

    @r2

    1

    + 16h31

    @3

    @r3

    1

    + O(h41) (3.4)

    as h1 ! 0, where we denote the value of a function at r = r1 by the subscript 1. Thesecond derivative @2 =@r2 can be expressed in terms of ; @ =@r; @2 =@2 at r = r1from (2.1) and the third derivative can be obtained by dierentiation of (2.1). The

  • 190 S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska

    third derivative involves @=@r and if we express this at r = r1 in the form

    @

    @r= (I 1)=h1 + O(h1); (3.5)

    where I denotes the grid point next to r = r1 along the normal, and neglect the errorterms in (3.4), (3.5), we nally obtain the approximation

    (1 h1=2r1)1() = (3=h21)[ I () f1() h1g1()] 12I ()(3=2r1)(1 2h1=3r1)g1() (3=2r21)(1 h1=r1)f001 () (h1=2r21)g001 (); (3.6)

    which is h2-accurate. This holds for all grid points on r = r1, and there is a similarformula for 2() at grid points on r = r2. This is obtained from (3.6) simply bychanging h1 to h1 and modifying the subscripts of 1; 1; f1; g1; r1 to 2; the subscriptI still refers to the grid point next to r2 along the normal direction on the fluid side.

    For the conditions on = we used exactly the same method of expanding alongthe inward normal, and nd

    (r) =3

    h22r2[ (r) I (r)] 12I (r) (3.7)

    for all grid points on = , where again I denotes the internal grid point nearest tothe boundary; equation (3.7) holds on both = , and r refers to the grid point underconsideration. This completes the numerical formulation of the problem. We have tosolve the nite-dierence equations (3.2), (3.3) for all internal grid points by some formof matrix inversion, and at the same time satisfy all the boundary conditions, whichmeans that equations (3.6) on r = r1; r2 and (3.7) on = must also be satised.

    An iterative method was used. An approximate solution of equations (3.2), (3.3)was obtained and then used to calculate the right-hand sides of (3.6), (3.7) in orderto estimate 1(); 2(); (r). Successive over-relaxation was used to solve equations(3.2), (3.3), and relaxation is also necessary to calculate the boundary values of from(3.6), (3.7). For both equations under-relaxation is necessary since the calculationsare very sensitive owing to the large factors multiplying some of the terms on theright-hand sides when h1; h2 are small. Thus if

    (c) denotes the calculated boundaryvalue of from equation (3.6) or (3.7) at any iteration (k) of the procedure, then thenext approximation to the boundary value was taken as

    (k+1) = (1 !)(k) + !(c); (3.8)where ! is a relaxation parameter. It is necessary to take ! 1 in order that theiterative procedure converges. The same equation (3.8) denes the over-relaxationprocedure which was applied to at each grid point in the numerical solution of(3.3), but here we can take ! > 1, in general. A similar equation was used for during solution of (3.2); here we can again take ! > 1 though not necessarily thesame as that applicable to the determination of from (3.3). In all cases the choice of! depends upon the Reynolds number and also, to some extent, upon the grid sizes.

    The whole iterative procedure was in fact carried out by performing one operationthrough the complete set of equations (3.3) followed by a complete iteration through(3.2). Then, boundary values were calculated according to equations (3.6), (3.7) byusing (3.8). This whole procedure was repeated until numerical convergence accordingto the criterion X

    n

    j(k+1)n (k)n j < 1;Xn

    j (k+1)n (k)n j < 2; (3.9)

  • Flow along a diverging channel 191

    where 1; 2 are assigned small parameters and n is the total number of grid points.The parameters 1; 2 depend upon the grid sizes. In the numerous solutions obtained,some of which are reported below, it was customary to obtain more accurate solutions,for given R; and boundary conditions, by successively halving the grid sizes in bothdirections. However, for consistency in applying the convergence criteria in (3.9),the parameters 1; 2 should be quadrupled on each occasion that the grid is halved.Furthermore, some degree of care is needed to ensure that numerical convergencehas taken place when the number of grid points is large. Thus, in the end, a visualinspection of the boundary values of vorticity was also made after a xed numberof iterations following the satisfaction of the tests (3.9), in order to ensure that nosignicant change took place.

    4. Numerical resultsUsing a scheme of the form described in x3, we solved the steady problem (2.1),

    (2.6), (2.7) and the set of boundary conditions (2.8). Also, it is convenient, in viewof the Jeery{Hamel similarity solution governed by (2.11) and (2.12), to make thechange of variable = r2 in (2.6) which makes independent of r in the similaritysolution and thus more readily approximated by nite dierences. This transformationmerely adds terms

    R1 4r

    @

    @r+

    2

    r2

    2 + R

    @

    @

    to the right-hand side of (2.6), when expressed in terms of . The changes which needto be made in the various equations of x3 and in the boundary conditions are fairlyroutine and are not given here.

    A number of trial solutions were rst carried out to test the numerical method, e.g.with the boundary conditions (2.8) given by

    f1() = 3(1 2=32)=2; f2() = =; g1() = g2() = 0; (4.1)and, in dimensionless variables, with r1 = 1; r2 = 7; = 1. (These values of r1; r2were used thereafter except where stated otherwise. They were chosen somewhatarbitrarily, but give an aspect ratio r2=r1 which is large enough to observe whethersome flows settle down to JH flows or not, small enough to prevent the need for veryexpensive calculations, and close enough to the value for the channel in the experimentof Nakayama (1988, gure 105).) Several Reynolds numbers were considered andthree grids were employed in each case, namely h1 = 0:25; h2 = 0:05 followed by thetwo grids obtained by successively halving each grid length. Use of h2-extrapolationenabled us to check the accuracy. For the lower Reynolds numbers it was possibleto use relaxation parameters ! of 1:4 or 1:5 for and with very rapid numericalconvergence, but these had to be gradually reduced as R was increased. The relaxationfactors appropriate to (3.6) and (3.7) were generally less than 0:5, reducing to 0:05for higher values of R. Care is necessary to ensure complete numerical convergencewhen ! is as low as 0:05. Finally, the numerical results of these tests indicated goodaccuracy with a very clear approach to limit solutions as h1; h2 ! 0.

    The method outlined in x3 was also modied to deal with boundary conditions ofthe form

    = f1(); = z1() at r = r1; = f2(); = z2() at r = r2: (4.2)

    Here too, in order to check the computer coding and to estimate the numerical errors,

  • 192 S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska

    we chose values of ; R for R2 > R > 0, and assigned to fi(); zi(); i = 1; 2, the formscorresponding to the JH solutions. It was then found that the calculated solutions were approximately independent of r. Various grids were used and the resultsfound to be consistent: with h2-extrapolation we obtained agreement to six decimalplaces with solutions of the ordinary dierential system (2.11), (2.12) found by usinga Runge{Kutta{Merson method.

    It will help, while interpreting the results below, to recall that a basic symmetricflow in the full sector 6 6 may be (i) stable to all small perturbations, (ii)unstable, yet stable to all small symmetric perturbations, or (iii) unstable to at leastone antisymmetric and one symmetric mode of perturbation. To highlight this, wehave made some numerical experiments in the half sector 0 6 6 with boundaryconditions ( = @2 =@2 = 0) at = 0 satised by all symmetric solutions over thefull sector. By reflection in = 0, the solution in the half sector gives a symmetricsolution in the full sector. The theory of linear spatial stability suggests that in case(i) R < R2() and flows in both the full and half sectors are stable; in case (ii)R2() < R < R3() and the flows in the full sector are unstable but flows in the halfsector are stable; and in case (iii) R3() < R and all JH flows in the full and halfsectors are unstable, and there exists neither a symmetric nor an asymmetric JH flowwith a single maximum of the velocity.

    With R = 0 we veried the predictions of perturbation theory which are describedbriefly at the end of x2. For = 1 we chose f1() = G(y; 1; 0); z1() = r21 G00(y; 1; 0)and f2() = G(y; 1; 0) + r

    l12 ^(y; 1; 0); z2() = r22 G00(y; 1; 0) rl122 ( ^00 + l21 ^); where

    l1(= 3:1841798 + 0:6890455i for = 1), ^, is the leading mode belonging to a+. (Note

    that 2(0) =12 > 1 = .) Since R = 0 the linearization gives the exact solution of

    the perturbed problem and the normalization of ^ is not important. In this way weveried the spatial decay of the disturbance at r2 as r decreases to r1. We repeatedthe calculation with the disturbance imposed at r = r1, but with the forms r

    l11 ^

    and r(l12)1 ( ^00 + l21 ^) added to f1() and z1() respectively. The mode here withl1(= 1:1841798 + 0:6890455i for = 1) is the leading mode for r increasing andbelongs to a. We again found that this disturbance decays as r increases to r2.We remark that some of these solutions are, of course, asymmetric. It is possible toimpose boundary conditions at both r1 and r2 by reference to (2.14), with = l1 orl1 depending on whether the disturbance decays as r decreases to r1 or increases tor2 respectively. This was done and our numerical results agreed well with the exactresult (2.14). We have not displayed any results concerning the pressure, but merelynote from equation (2.13) that the Stokes pressure (i.e. Rp in the limit as R ! 0) isinversely proportional to the square of the distance downstream.

    We conclude plausibly that the numerical method and the coding of the problem aresatisfactory. It should be noted, however, that we could not evaluate the eigenvaluesfrom the numerical solutions of the partial dierential problem. We presume that thisis because several modes are superposed in the numerical solutions, the modes decaytoo slowly (their decay being algebraic) and the sector is not long enough (with ratior2=r1 only 7) for the leading mode to dominate the others anywhere.

    Before describing our results for R 6= 0, we describe those for R = 0 and 2 < < 3,where R2(2) = 0; R3(3) = 0; in fact 2 =

    12; 3 2:246 when R = 0. The value

    = 1:7 was chosen as typical to ensure in this way that the JH flow (of type II2) inthe full sector is spatially unstable to the l1 and l1 antisymmetric modes, althoughthe JH flow in a half sector is spatially stable to all modes (2.14).

    For boundary conditions, we rst imposed the appropriate JH solution (of type

  • Flow along a diverging channel 193

    3

    2

    1

    02 4 6 7

    r

    Full

    Half

    Half andfull

    (r, 12)

    1

    Figure 4. Values of (r; 12) against r for R = 0; = 1:7 for flows in the full and half sectors:

    solid curve for the perturbation imposed at r = r2 (the results for the full and half sectors areindistinguishable); broken curve for the half sector with the perturbation imposed at r = r1; chainedcurve for the full sector with the perturbation imposed at r = r1.

    I; II1 or II2) at both r = r1 and r2; the results for both full and half sectors were thesame and the solutions were approximately independent of r. Even though the flowis predicted to be spatially unstable in the full sector, the calculation failed to revealthis instability. So we next investigated the eect of perturbations at r = r1 and r2in turn: we added (2 2)24 (i.e. an antisymmetric disturbance) to the basic JHstreamfunction. We found that, irrespective of whether the problem is posed in thehalf or full sector, when the disturbance is imposed at r = r2 the approach to theJH solution at r = r1 is monotonic, whereas if the disturbance is imposed at r = r1there is an overshoot in the approach to the JH solution at r = r2. Further, when thedisturbance is imposed at r = r2 there is very little dierence between the full- andhalf-sector results, but there is a signicant dierence when the disturbance is imposedat r = r1. We presume that this dierence is due to the relative size of the two leadingantisymmetric eigenvalues from a+ and

    a (= 1:85891 and 0:14109 respectively) andthe shortness of the channel (r2=r1 = 7). The plots of (r;

    12) in gure 4 illustrate

    these properties; note that the results for the half and full sectors clearly dier whenthe JH flow is perturbed at r = r1. We anticipate that the overshoot in the full sectorbecomes greater as the radius ratio is increased: we conrm this later for some othervalues of and R when the ratio is increased.

    We next present results for R > 0. We chose = 1 and sought solutions in boththe full and half sectors with 2:9 = R2(1) < R < R3(1) = 4:4: we chose R = 3:5 astypical. It will be noted that, as for R = 0; = 1:7, the JH flow in a full sector isspatially unstable while that in a half-sector is stable to the modes (2.14). We startedthe calculations by imposing the JH solution at r = r1 and r2 in both the full- andhalf-sector congurations, and both sets of results agree with the known JH solutionto about 5 signicant gures: the flow predicted is again independent of r. This isthe same as when R = 0; = 1:7. Indeed, even when the streamfunction at r = r1; r2is perturbed by the addition of (2 2)24, the results are still similar to those forR = 0; = 1:7. However, we proceeded further in this case by increasing the radiusratio r2=r1 of the sector from 7 to 13 and nally to 25. With the perturbation imposedat r = r2, the results for the full sector diered very little (about 0:02%): note that

  • 194 S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska

    r251370 1

    0.88

    0.90

    0.92

    0.94

    251370 1

    0.90

    0.91

    0.92

    0.93

    r

    (a) (b)

    (r, 12) (r, 12)

    Figure 5. Values of (r; 12) against r for R = 3:5; = 1 and various ratios r2=r1. (For unperturbed

    JH flow the curve would be the horizontal line (r; 12) = 0:89857.) (a) The flow in the half sector,

    with the perturbation imposed at r = r1. The solid curve denotes the results for r2=r1 = 25, forr2=r1 = 13, for r2=r1 = 7. (b) The flow in the full sector, with the perturbation imposed at r = r1.The solid curve denotes the results for r2=r1 = 25, for r2=r1 = 13, for r2=r1 = 7. The analogousresults for the half sector with r2=r1, shown by a broken curve are repeated for comparison.

    for this comparison it was necessary to re-scale the interval r1 < r < r2. Further notethat, as reported for R = 0; = 1:7, the dierence between the full- and half-sectorsolutions was small (typically about 0:1%) when the perturbation was imposed atr = r2. With the perturbation imposed at r = r1, however, the dierence between thefull- and half-sector solutions increases as the radius ratio is increased: for the halfsector, increasing the ratio changes the results very little, but for the full sector itchanges them substantially. The plots of (r; 1

    2) in gure 5 illustrate these properties,

    showing the eect of lengthening both the half and full sectors. Note that lengtheningthe full sector increases the overshoot. All this conrms that the basic JH flow in thehalf sector is stable but that in the full sector is not.

    Consider next values of ; R at which there is no JH solution with uniformlyoutward flow, i.e. with radial velocity u > 0 for all . We chose = 1 again, withR = 7:5 (> R3(1) = 4:4), for our calculations. Because there is no uniformly outwardJH solution on which to base the boundary conditions, we imposed at r = r1; r2 theStokes solution for = 1; R = 0, namely

    G(y; 1; 0) =sin 2y 2y cos 2sin 2 2 cos 2 ;

    together with the vorticity associated with this streamfunction.We found that the results for the full and half sectors are the same: the flow in

    the full sector is symmetric and coincides with the flow found for the half sector.However, adding (1 2)24 to the streamfunction at the inlet r = r1 in order toperturb the flow asymmetrically, we found that the flow in the full sector is changedmuch more than the flow in the half sector. Similarly, adding the same perturbation

  • Flow along a diverging channel 195

    1.0

    0.9

    0.8

    75310

    0.7

    0.8

    0.9

    Fullsector

    r/r1

    Halfsector

    Figure 6. Values of (r; 12) against r for R = 7:5; = 1. The values for flow in the full sector

    with no perturbation are shown by a full curve, those for flow in the half sector, also with noperturbation, by a broken curve. The corresponding cases with the perturbation imposed at r = r1are shown by , and with the perturbation imposed at r = r2 are shown for the full sector by +and for the half sector by . Note the displaced scales for the two distinct cases.

    at the outlet, we found that the flow in the full sector is changed more than theflow in the half sector, but the dierence is much greater when the inlet condition ismodied. The plots of (r; 1

    2) in gure 6 illustrate these properties. Note that, for a

    flow in either the half or the full sector, the results when the perturbation is imposedat r = r1 cross those of the unperturbed flow. All these results suggest that the flowin the full sector is unstable but the flow in the half sector is stable, although morecalculations with larger radius ratios are needed to substantiate this suggestion. Alsowe presume that the eigenvalue belonging to the leading mode in a+ has a greaterreal part than that of the eigenvalue belonging to the leading mode in a, althoughboth eigenvalues are unknown (we assume that, even though there is no JH flow ofphysically acceptable form, there are the usual two families of eigensolutions).

    For both a comparison of theory and experiment and an example of flow in thefull sector with R R3(), we sought to simulate the experiment of Nakayama (1988,gure 105). He drove a flow through a channel with = =18; R = 300; we havebeen unable to ascertain the inlet and outlet conditions, but estimate that the aspectratio r2=r1 5. It appears that his flow was very smooth and steady. Note thatthe hydrogen-bubble proles in his photograph are not quite the same as velocityproles at given radii. Also the observed asymmetric flow (see our gure 1) with onemaximum and two minima of the velocity does not correspond to any JH flow; it isplausibly an asymmetric steady flow following a supercritical pitchfork bifurcation ata lower value of the Reynolds number. We have obtained many results for flow inboth the full and the half sectors with = =18; r2=r1 = 7 and values of R increasingfrom 5. In the absence of any JH flow for R > R3, we chose as boundary conditions (r; ) = G(y; ; 0) for 6 6 at r = r1; r2, where G is the Stokes solution togetherwith the associated conditions for the vorticity, and then, in separate calculations,

  • 196 S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska

    (a) (b)

    u840

    1

    1

    y

    Figure 7. Results for R = 5; = =18. (a) Streamlines and vorticity contours. The streamlines, for = n=10; n = 0; 1; 2; :::; 10, are shown in the lower-half sector and vorticity contours in the upper.(b) Velocity proles (u versus y) at r = 1; 2; 3; 4; 5; 6; 7. (They can be identied by noting that thevalue of u on the centreline decreases with increasing r.) Note that those proles corresponding tor = 1; 7 are imposed as boundary conditions.

    added the perturbations (1 (=)2)2(=)4 at r = r1 or r2. Next we describe ourresults with the aid of graphs.

    When R = 5 the results for the half and full sectors are indistinguishable. Theyare, moreover, indistinguishable whether the perturbation is imposed at r = r1 orr2; so gure 7(a) shows the streamlines in the lower-half sector and the contoursof constant vorticity in the upper-half sector of the symmetric flow. Note that thestreamlines are fairly straight and the vorticity is roughly constant in a large areaof the sector. Also the velocity proles at r = 1; 2; 3; 4; 5; 6; 7 are shown in gure7(b) (but remember that the proles at r = 1; 7 are given directly by the boundaryconditions). Note that u is approximately inversely proportional to r (as expectedbecause R < R2(=18) = 28:68; R3(=18) = 31:03). We have found the variation ofpressure with r along = 0;=4;=2;3=4;, and we show the results in gure8(a). It will be noted that we have plotted p(r; ) p(1; ) on each ray so that thepressure drop is clearly evident although the variation is not of inverse-square form.

    Similar patterns were found for R = 10; 25. However, at R = 25 the vorticitycontours are appreciably dierent, as demonstrated in gure 9, which shows thevorticity contours for the flow in the full sector with the perturbation imposed at theinlet r = r1 (detailed investigation of the streamlines near the inlet, shown in gure10(a), reveals the eect of the perturbation). Note that the flow is nearly symmetric.The velocity proles at r = 1; 2; 3; 4; 5; 6; 7 are shown in gure 10(b) (again the eectof the perturbation is revealed by the gure). Further, the pressure varies with rdierently in that it increases for some values of . This is clearly shown in gure 8(b)where we display the pressure variation versus r for = 0;=4;=2;3=4;.

    When R = 50, the flow in the full sector is again nearly symmetric, but the symmetrycan be strikingly broken when a small perturbation is imposed: there may be an eddywith reversed flow near one of the walls. Figure 11 shows the streamlines when aperturbation is imposed at r = r2 and gure 12(a) when it is at r = r1. The occurrenceof a closed streamline implies that there is separation and reattachment on the lowerwall. Of course, taking the mirror image of the perturbation in the centreline wouldresult in the eddy being near the other wall. The eects of the perturbation at r = r2are signicant only locally. The asymmetry of the streamlines in gure 12(a) canalso be seen in a plot of the vortex contours (not shown here). The velocity prolesare shown in gure 12(b). When the flow is symmetric and when the perturbationis imposed at r = r2, the pressure variations along the channel in both cases aresimilar to that shown in gure 8(b), with any asymmetry conned to the immediatevicinity of r = r2. However when the perturbation is imposed at r = r1 the results do

  • Flow along a diverging channel 197

    (a) (b)0

    2

    4

    Dp

    1 3 5 7r

    h=

    34

    140

    12

    0

    12

    h=

    0

    1

    1

    Dp

    (c)

    0

    0.4

    0.4

    Dp

    1 3 5 7r

    0.8

    12

    h=

    0

    12

    12

    1 3 5 7r

    0

    1

    Dp

    h=34

    34

    0

    1 3 5 7r

    (d)

    14

    Figure 8. Variation of pressure, p = p(r; ) p(1; ), along the channel, at indicated values of for (a) half sector, R = 5, (b) full sector, R = 25, (c) full sector, R = 50 and (d) half sector, R = 300.

    Figure 9. Vorticity contours for R = 25; = =18 for flow in the full sector with the perturbationimposed at r = r1.

    (a) (b)

    u840

    1

    1

    y

    Figure 10. Results for R = 25; = =18 for flow in the full sector with the perturbation imposedat r = r1. (a) Streamlines for = 1 + 2n=19 with n = 0; 1; 2; :::; 19. (Note that the centreline hasbeen shown to indicate the asymmetry.) (b) Velocity proles at r = 1; 2; 3; 4; 5; 6; 7. (They can beidentied by noting that the value of u on the centreline decreases with increasing r.)

  • 198 S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska

    Figure 11. Streamlines for R = 50; = =18 for flow in the full sector with the perturbationimposed at r = r2; they are shown for = 1 + n=5 with n = 0; 1; 2; :::; 10.

    (a)(b)

    u620

    1

    1

    y 7 6 5 43 2 1

    Figure 12. Results for R = 50; = =18 for flow in the full sector with the perturbation imposedat r = r1. (a) Streamlines for = 1 0:243n with n = 0; 1; 2; :::; 9 and = 1:43 (the minimum)indicated as a point. (b) Velocity proles at various values of r (as indicated).

    vary signicantly with (as might be expected), and those for = 0;=2; aredisplayed in gure 8(c).

    The solution for R = 50 for the full sector did not converge readily and requireda large amount of computing time, so we did not proceed to larger values ofR. However, for the half sector the calculations are easier, and results have beenobtained up to R = 300.

    When R = 100, the eddy is found even for the flow in a half sector, and againthe perturbed flows due to imposing the perturbation at the inlet and the outletare indistinguishable. Figure 13(a) shows the streamlines and the vorticity contours.Figure 13(b) shows the velocity proles; note that there is a large re-adjustmentnear the outlet, where the boundary condition is imposed. Renement of the gridis progressively more important as R increases: for example, the undulations andislands in the vorticity contours near the right-hand end of the upper half of gure13(a) are presumably due to the coarseness of the grid (which may lead to errors inthe calculations or the graph plotting) because these features are exaggerated in ourresults (not shown here) with a coarser grid. However, the streamlines are much lessaected by coarsening of the grid.

    Finally, when R = 300, the flow in the half sector has an eddy; and the unper-turbed flow in the full sector is symmetric with an eddy near each wall which isindistinguishable from an eddy in the flow in a half sector. Figure 14(a) shows thestreamlines and vorticity contours, and gure 14(b) the velocity proles. Again thereis a large re-adjustment of the flow near the outlet (a consequence of the concentra-tion of streamlines near the outlet shown in gure 14a). In gure 8(d) we show thepressure variation as a function of r for = 0;=4;3=4;. The rapid variationof the pressure near the exit is presumably a consequence of the concentration of thestreamlines there.

  • Flow along a diverging channel 199

    (a) (b)

    u820

    1

    1

    y

    Figure 13. Results for R = 100; = =18 for flow in the half sector. (a) Streamlines and vorticitycontours. The streamlines, for = 0:138n with n = 0; 1; 2; :::; 9 and = 1:378 (the minimum)indicated as a point, are shown in the lower-half sector and vorticity contours in the upper. (b)Velocity proles at r = 1; 2; 3; 4; 5; 6; 7. (They can be identied by noting that the value of u on thecentreline decreases with increasing r.)

    (a) (b)

    u0

    1

    1

    y

    2

    Figure 14. Results for R = 300; = =18 for flow in the half sector. (a) Streamlines and vorticitycontours. The streamlines, for = 0:153n with n = 0; 1; 2; :::; 9 and = 1:53 (the minimum)indicated as a point, are shown in the lower-half sector and vorticity contours in the upper. (b)Velocity proles at r = 1; 2; 3; 4; 5; 6; 7. (They can be identied by noting that the value of u on thecentreline decreases with increasing r.)

    The flow patterns found for R = 50; 100 and 300 have led us to hazard thefollowing conjecture. As R increases, the asymmetry of the flow in the full sectorbecomes more pronounced, the eddy getting larger and the rest of the flow beingmore narrowly conned to a thin layer near the wall on the other side, at = ingure 12(a). The thinness of the layer suggests that the bulk of the asymmetric flowmay be calculated approximately, for large R, by allowing a slip velocity at = ;for convenience we suggest a symmetry condition precisely the same as the conditionapplied above at = 0 for flow in a half sector { but applied now at = , of course.Thus the calculation of the flow in the half sector for semi-angle 2 is used to ndapproximately the eddy for the asymmetric flow in the full sector with semi-angle .In addition, the description of the flow in the full sector with semi-angle might becompleted by solving the boundary-layer problem near = driven by the outer flow,i.e. the slip velocity found from the problem for the half sector with semi-angle 2.

    The choice, in a theory, of boundary conditions at the inlet and outlet of a divergingsector to represent an experiment faithfully is usually dicult. Experiments are oftenconducted with a channel which has an inlet consisting of parallel walls long enoughto establish a parabolic velocity prole, some divergent test section, and an outletthat is incompletely described.

    Often in computational fluid dynamics a parabolic prole is imposed at the inlet(say, condition (2.4) with f1() = 3(1 2=32)=2; g1() = 0) and longitudinalflow at the outlet (say, condition (2.5) with appropriate f2 and g2() = 0). Now animplication of the results reported above is that if R < R2(), with a sector suitablylong (depending on the size of the most signicant eigenvalue), then the particular

  • 200 S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska

    choice of inlet and outlet conditions aects the flow little in the middle region of thesector, but if R > R2() then the choice aects the flow substantially irrespectively ofthe length of the sector.

    5. ConclusionsIn addition to verifying a previous conjecture about the spatial development of

    the eigenmodes for R = 0, we also evaluated our computer programs by using theanalytical results of JH flows as a benchmark, before we embarked on the moreinteresting numerical experiments. This gives condence in the reliability of ournumerical results.

    The influence of the inlet and outlet of the channel has been found to be far-reaching even when the basic JH flow is spatially stable, because algebraic decayis much slower than the exponential decay which is characteristic of parallel flows,the exponential decay to which many have become accustomed. When the basicflow is spatially unstable the conditions at the inlet and outlet influence the wholeflow in the channel, and it may be said that Saint-Venants principle is invalid. Ithas been seen that, on the whole, the inlet conditions aect the flow more stronglyand further into the interior than do the outlet conditions, whether R < R2() orR > R2(). It is plausible that this relative importance of the inlet conditions is due tothe convection of vorticity, but it seems to be a quantitative rather than a qualitativeproperty mathematically.

    The spatial modes of the linearized problem have provided a fruitful means tointerpret the influence of the inlet and outlet conditions on a stable flow in thediverging channel, but it should be noted that the linearized problem is self-adjointonly when the Reynolds number is zero, and therefore that the development in spaceof two or more superposed modes may not be easy to predict. Indeed, by analogywith the temporal problem of superposing stable modes (cf. Gustavsson 1991; Butler& Farrell 1992; Reddy & Henningson 1993), a small perturbation of a stable JHflow at the outlet and inlet may lead to a very large perturbation of the flow in theinterior of the channel when the Reynolds number is large; however, we have notobserved a very strong transient spatial amplication, perhaps because most of ourcalculations have been made for moderate values of the Reynolds number.

    Note that the numerical solutions for the full sector show that when R = 25 theflow is nearly symmetric whereas when R = 50 it is clearly asymmetric. This impliesthat there is a pitchfork bifurcation at an intermediate value of R. The value ofR2(=18) = 28:68 for the subcritical pitchfork bifurcation of JH flows is a plausibleestimate for the value of R at the supercritical pitchfork bifurcation of this forcedflow in a sector of radius ratio 7.

    It has been impossible for us to simulate exactly the experiment of Nakayama (1988,gure 105) because of our ignorance of both his quantitative results and the preciseconditions at the inlet and outlet of his channel, and we have shown how importantthose conditions are when R > R2(). Nonetheless, our attempted simulation of theexperiment has been successful on the whole, even though we could not compute theasymmetric flow accurately at as large a value of the Reynolds number as 300 atwhich Nakayama conducted his experiment, and the calculated eddies for symmetricflow at R = 300 and for asymmetric flow at R = 50 are larger than those observed inthe experiment.

    Our choice of the special channel in a sector for intensive study was motivatedby the need to use a prototype with which analytic and asymptotic results could be

  • Flow along a diverging channel 201

    used to interpret results of computational fluid dynamics, rather than the intrinsicengineering importance of the flow in a sector. However, we believe that the set offlows in a sector with various boundary conditions is qualitatively similar to the set offlows in a wide class of diverging channels used as diusers, although the particularboundary conditions (2.5) are a poor model of the outlet of a diuser.

    This paper describes only two-dimensional flows in symmetric channels. Thesymmetry is, of course, only an idealization dicult to realize in an algorithm ofcomputational fluid dynamics and impossible to realize in the laboratory. So thetheory of imperfections is important in the practical interpretation of our results (cf.Sobey & Drazin 1986; Fearn et al. 1990). Also all laboratory experiments of flowin two-dimensional channels are, of course, of three-dimensional flow. If the cross-section of the channel at each station is a rectangle, say, perhaps with a large aspectratio to simulate a two-dimensional flow, then we may expect not only the symmetrybreaking of a pitchfork bifurcation as described above, but also a pitchfork or otherbifurcation to describe the symmetry breaking when a strongly three-dimensional flowrst appears as the Reynolds number slowly increases.

    We have treated some bifurcations of flows in a channel rather than the later stagesof transition to turbulence because it is desirable to sort out the fundamentals beforethe more interesting, more challenging and more useful problems of flow at largevalues of the Reynolds number. Nonetheless, let us nally note King & Stewarts(1991) conjecture that, as the Reynolds number R of flows in a given symmetricconguration slowly increases, rst symmetry is broken and then it is recovered asthe asymmetric attractors increase in size in phase space until they collide. Thisconjecture, if appropriately modied, seems as plausible when the conguration ofthe flows is asymmetric as when it is symmetric, because more than one attractor mayoccur at a given value of R and the domains of attraction in phase space increase insize as R becomes large. So, even though there is no exact symmetry of a channel,we may expect the flow to bifurcate and rst get further from, and eventually closerto, a symmetric flow as the Reynolds number increases. One might go further andhypothesize that there is a unique attractor for flow in a given channel as R tendsto innity. There is no strong argument to justify these hypotheses but they do seemplausible descriptions of observations of flows in laboratory channels.

    We are grateful to Dr I. B. Stewart for help with computer plots.

    Note added in proof. Some may benet by reading the interesting recent paper byTutty (1996), which reports computations of steady flows in a very long channel withplane walls over most of its length.

    REFERENCES

    Banks, W.H.H., Drazin, P.G. & Zaturska, M. B. 1988 On perturbations of Jeery{Hamel flow.J. Fluid Mech. 186, 559{581.

    Batchelor, G.K. 1967 An Introduction to Fluid Dynamics . Cambridge University Press.

    Borgas, M. S. & Pedley, T. J. 1990 Non-uniqueness and bifurcation in annular and planar channelflows. J. Fluid Mech. 214, 229{250.

    Buhler, K. & Kruckels, J. 1990 Stromungsformen mit Ablosung im ebenen Diusor. Z. Angew.Math. Mech. 70, T475{477.

    Butler, K.M. & Farrell, B. F. 1992 Three-dimensional optimal perturbations in viscous shearflows. Phys. Fluids A 4, 1637{1650.

  • 202 S. C. R. Dennis, W. H. H. Banks, P. G. Drazin and M. B. Zaturska

    Cliffe, K.A. & Greenfield, A. C. 1982 Some comments on laminar flow in symmetric two-dimensional channels. Rep. TP 939. AERE, Harwell.

    Dean, W.R. 1934 Note on the divergent flow of fluid. Phil. Mag. (7) 18, 759{777.

    Drazin, P.G. & Reid, W.H. 1981 Hydrodynamic Stability . Cambridge University Press.

    Fearn, R.M., Mullin, T. & Cliffe, K.A. 1990 Nonlinear flow phenomena in a symmetric suddenexpansion. J. Fluid Mech. 211, 595{608.

    Fraenkel, L. E. 1962 Laminar flow in symmetrical channels with slightly curved walls. I. On theJeery{Hamel solutions for flow between plane walls. Proc. R. Soc. Lond. A 267, 119{138.

    Georgiou, G.A. & Eagles, P.M. 1985 The stability of flows in channels with small wall curvature.J. Fluid Mech. 159, 259{287.

    Goldshtik, M., Hussain, F. & Shtern, V. 1991 Symmetry breaking in vortex-source and Jeery{Hamel flows. J. Fluid Mech. 232, 521{566.

    Gustavsson, L.H. 1991 Energy growth of three-dimensional disturbances in plane Poiseuille flow.J. Fluid Mech. 224, 241{260.

    Hamadiche, M., Scott, J. & Jeandel, D. 1994 Temporal stability of Jeery{Hamel flow. J. FluidMech. 268, 71{88.

    Hooper, A. P., Duffy, B. R. & Moffatt, H.K. 1982 Flow of fluid of non-uniform viscosity inconverging and diverging channels. J. Fluid Mech. 117, 283{304.

    King, G. P. & Stewart, I. N. 1991 Symmetric chaos. In Nonlinear Equations in the Applied Sciences,(ed. W. F. Ames & C. F. Rogers), pp. 257{315. Academic.

    Nakayama, Y. (Ed.) 1988 Visualized Flow. Pergamon.

    Reddy, S. C. & Henningson, D. S. 1993 Energy growth in viscous channel flows. J. Fluid Mech.252, 209{238.

    Sobey, I. J. & Drazin, P.G. 1986 Bifurcations of two-dimensional channel flows. J. Fluid Mech. 171,263{287.

    Sobey, I. J. & Mullin, T. 1992 Calculation of multiple solutions for the two-dimensional Navier{Stokes equations. Proc. ICFD Conference, Reading.

    Tutty, O.R. 1996 Nonlinear development of flow in channels with non-parallel walls. J. Fluid Mech.326, 265{284.

    Woods, L. C. 1954 A note on the numerical solution of fourth order dierential equations. Aero. Q.5, 176{184.


Recommended