Sunteți pe pagina 1din 58

Batchelor versus Stewartson ow structures in a rotor-stator cavity with throughow

Sebastien Poncet, Marie-Pierre Chauve, and Roland Schiestel


I.R.P.H.E., UMR 6594 CNRS-Universit e dAix-Marseille I & II Technop ole Ch ateau-Gombert, 49 rue F. Joliot-Curie, 13384 Marseille c edex 13 - FRANCE - Fax. 33 (4) 96 13 97 09 (Dated: 20th April 2005) The present work considers the turbulent ow inside a high speed rotor-stator cavity with or without superimposed throughow. New extensive measurements made at IRPHE by a two component laser Doppler anemometer technique and by pressure transducers are compared to numerical predictions based on one-point statistical modeling using a low Reynolds number second-order full stress transport closure (RSM model). The advanced second-order model provides good predictions for the mean ow as well as for the turbulent eld and so is the adequate level of closure to describe such complex ows. A better insight into the dynamics of such ows is also gained from this study. Indeed the transition between a Batchelor type of ow with two boundary layers separated by a central rotating core and a Stewartson type of ow with only one boundary layer on the rotating disk is characterized in the (r , Ro) plane, where r is the dimensionless radial location and Ro a modied Rossby number. The 5/7 power-law of Poncet et al. [1] describing the mean centripetal ow in a rotor-stator system is extended to dierent aspect ratios and to the case of centrifugal Batchelor type of ows.

Electronic address: poncet@irphe.univ-mrs.fr,Tel.33(4)96.13.97.75 Electronic address: chauve@irphe.univ-mrs.fr,Tel.33(4)96.13.97.76 Electronic address: schieste@irphe.univ-mrs.fr,Tel.33(4)96.13.97.65

2
I. INTRODUCTION

The interest in rotating disk ows is multiple and has justied many works since the pioneering work of Ekman [2] in 1905. It has a major interest in turbomachinery, but examples where rotation and throughow are associated can also be found in oceanography, geophysics or astrophysics. From a fundamental point of view, the rotor-stator problem is also one of the simplest conguration where rotation inuences turbulence eld and where exact solutions of the Navier-Stokes equations can be found for laminar ows. Von K arm an [3] studied the laminar ow over an innite rotating disk in a quiescent uid. He simplied the equations of motion governing the ow on the free disk to nonlinear dierential equations using the assumption of axisymmetry. He showed that the ow is conned in a thin boundary layer on the disk. He studied also the turbulent case by using momentum integral methods with power law velocity proles. The Von K arm an analysis was subsequently followed by B odewadt [4] who investigated numerically the ow over a innite stationary plane with an outer ow in solid body rotation. Batchelor [5] solved the system of dierential equations relative to the stationary axisymmetric ow between two disks of innite radius. He specied the formation of a non-viscous core in solid body rotation, conned between the two boundary layers which develop on the disks. From 1953 to 1983, this division of the ow into three distinct zones was the subject of an intense controversy: Stewartson [6] found indeed, in 1953, that the tangential velocity of the uid can be zero everywhere apart from the rotor boundary layer. The problem of the existence or not of a core in solid body rotation justied many works until Kreiss and Parter [7] proved the existence of a multiple class of solutions discovered numerically in fact by Mellor et al. [8]. Zandbergen and Dijkstra [9] showed also that the similarity equations do not generally have unique solutions and that the two solutions advocated by Batchelor and Stewartson can thus both be found from the similarity solutions. In the case of a laminar ow between two nite disks, Brady and Durlofsky [10] found that ows in an enclosed cavity are of Batchelor type, while open-end ows are of Stewartson type. Daily and Nece [11] performed experimental and theoretical studies in a closed rotorstator cavity. They pointed out the existence of four ow regimes according to the Reynolds
2 / and to the aspect ratio G = h/R2 ( is the angular velocity of the number Re = R2

rotating disk, R2 the outer rotating disk radius, the kinematic viscosity and h the inter-

3 disk space): two laminar and two turbulent regimes, each of which corresponding either to merged or separated boundary layers. This classication was conrmed numerically by Lance and Rogers [12] and Owen and Rogers [13]. Daily et al. [14] measured the average velocity proles and put forward the importance of the outow on the development of the Ekman layer and on the values of the entrainment coecient of the uid. Firouzian et al. [15] performed velocity and pressure measurements for a wide range of ow rates and rotational speeds. They tested also the eect of inlet geometry on the ow between two disks enclosed by a peripheral shroud. Owen and Rogers [13] showed experimentally the dominating inuence of a centripetal ow in the determination of the entrainment coecient K of the uid (K is the ratio between the mean tangential velocity in the central rotating core and that of the disk at the same radius). The experiment of Itoh et al. [16] provided a great contribution to the understanding of the turbulent ow in a shrouded rotor-stator system. The authors measured the mean ow and all the Reynolds stress components and brought out the existence of a relaminarized region even for high rotation rates. Cheah et al. [17] performed detailed measurements of mean velocity and of the turbulent ow eld inside a rotor-stator system. Debuchy et al. [18] studied the radial inow in a rotor-stator system. They found analytical solutions and provided experimental data in good agreement with the features of their asymptotic model. More recently, Poncet et al. [1] measured the entrainment coecient of the uid K for dierent aspect ratios in the case of a centripetal throughow and showed that K depends on a local ow rate coecient Cqr according to a 5/7 power law which has been also determined analytically. Beside the theoretical aspect, turbulent rotating disk ows can be considered as useful benchmarks for numerical simulations because of the numerous complexities embodied by this ow. Chew [19] was the rst to study numerically the ow inside a rotor-stator cavity with centrifugal throughow using a low Reynolds number k model. Chew and Vaughan [20] studied this type of ow with and without a superimposed throughow with a model based on a mixing length hypothesis inside the whole cavity. Their results were quite comparable to the experimental data of Daily and Nece [11] and Daily et al. [14] apart from a relaminarization area close to the rotating axis. The model of Iacovides and Theofanopoulos [21] used an algebraic modeling of the Reynolds stress tensor in the fully developed turbulence area and a mixing length hypothesis near the wall. It provided good results in the case of a rotor-stator ow with and without throughow but some discrepancies remain on the

4 Ekman layer thickness and the rotation rate in the central area. Iacovides and Toumpanakis [22] tested four turbulence models and showed especially that the Reynolds stress model was an appropriate level of closure to describe rotor-stator ows in a closed cavity. Schiestel et al. [23] have used both a low-Reynolds number k model near the walls and an Algebraic Stress Model (ASM) in the core of the ow. Second order informations were also found to be necessary in turbulence closure to get a sucient degree of universality in predicting highly rotating ows. Debuchy [24] realized a comparative study between experimental results relative to the ow inside a rotor-stator with a superimposed centripetal throughow and the computations obtained with a numerical model developed from an asymptotic approach. But the limitations inherent to the turbulence models and to the representations of the boundary conditions, did not allow to carry out reliable predictions. Later Elena and Schiestel [25] proposed also some numerical calculations of rotating ows based on a zonal approach. They have also used a new modeling of the Reynolds stress tensors derived from the Launder and Tselepidakis [26] one. It provides a better prediction than the simpler model of Hanjalic and Launder [27]. But there also, the authors emphasized the too high laminarization of the ow in comparison with the expected results. Iacovides et al. [28] tested two low Reynolds number models: a classical k model and a modied Reynolds stress model (RSM model) which takes into account the eect of the rotation. Poncet et al. [29] compared pressure and velocities measurements with numerical predictions based on an improved version of the Reynolds stress modeling of Elena and Schiestel [30] for centripetal and centrifugal throughows [29]. All the comparisons were in excellent agreement for the mean and turbulent elds. The RSM model is a valuable tool to describe the turbulent ow with or without a superimposed throughow compared to the DNS which are limited at present time to small rotation rates. Since the pioneering contribution of Fromm [31], all the DNS are restricted to simulate turbulent rotor-stator ows with Reynolds number not higher than 3 4 105 [3234]. The present paper is devoted to the study of the turbulent ow in a rotor-stator system of large aspect ratio (regime IV [11]) when a throughow is superimposed on the rotating uid. The basic ow belongs to the Batchelor type family: the two boundary layers are separated by a central rotating core. By superimposing a centripetal ow, the core rotates faster than in the closed cavity but the ow remains of Batchelor type. On the contrary, by superimposing a centrifugal ow, the ow gets a Stewartson type of ow with only one

5 boundary layer on the rotating disk and a quasi zero tangential velocity outside. The purpose of this work is to compare advanced turbulence modeling to the new data obtained from extensive velocity and pressure measurements when a throughow is superimposed and to characterize the transition between Batchelor and Stewartson ow structures.

II.

EXPERIMENTAL SET-UP A. Apparatus

The cavity sketched in gure 1 is composed of a smooth stationary disk (the stator) and a smooth rotating disk (the rotor). A xed shroud encloses the cavity. The rotor and the central hub attached to it rotate at the uniform angular velocity . The mean ow is mainly governed by three control parameters: the aspect ratio of the cavity G, the rotational Reynolds number Re based on the outer radius of the rotating disk and the ow rate coecient Cw dened as follows: G= h R2 Re =
2 R2

Cw =

Q R2

where is the kinematic viscosity of water, R1 , R2 the inner and outer radii of the rotating disk, R3 the outer radius of the cavity, R4 its central opening and Q the superimposed throughow. Cw = 0 corresponds to a closed cavity. Cw > 0 (resp. < 0) denotes the case where a centripetal (resp. centrifugal) throughow is superimposed. The interdisk space h can vary between 3 and 12 mm (0.012 G 0.048). In the present work, the interdisk space h and the radial gap e = R3 R2 are respectively xed to 9 (G = 0.036) and 3 mm. A pump allows to impose a variable throughow Q. The measurement of the ow rate is performed by an electromagnetic ow-meter, located at the exit of the cavity. The rotation of the disk is produced by a 5.5 kW electric servo-motor. A variable speed numerical controller directs the angular velocity . The accuracy on the measurement of the angular velocity and on the throughow is better than 1%. In the case of centripetal throughows, the uid is entrained in prerotation. As the disk rotates, water is sucked through the central opening of the cavity situated above the hub. The incoming uid is entrained in rotation while passing through a breakthrough crown mounted underneath the rotor and linked to it. It enables to increase the tangential velocity

6 of the uid, and consequently limits the inuence of the non rotating cylindrical wall. This prerotation is achieved through 48 holes having a diameter of 10 mm calibrated in order to entrain enough uid when the disk rotates. According to the superimposed centripetal throughow, the mean prerotation rate ranges between 0.43 and 0.54 [35]. In order to avoid cavitation eects, the cavity is maintained at rest at a pressure of 2 bar. Pressurization is ensured by a tank-buer and is controlled by two pressure gauges. The temperature is maintained constant (23 C ) using a heat exchanger which allows to remove the heat produced by friction in order to keep constant the density and the kinematic viscosity of water.

B.

Instrumentation and measurements

The measurements are performed using a two component laser Doppler anemometer (LDA) and also using pressure transducers. The LDA technique is used to measure from above the stator the mean radial Vr and tangential V velocities as well as the associated
Reynolds stress tensor components Rrr = vr v /(r)2 , R = v2 /(r)2 in a = vr2 /(r)2 , Rr

vertical plane (r, z ). This method is based on the accurate measurement of the Doppler shift of laser light scattered by small particles (Optimage PIV Seeding Powder, 30 m) carried along with the uid. Its main qualities are its non intrusive nature and its robustness. About 5000 validated data are necessary to obtain the statistical convergence of the measurements. It can be noticed that, for small values of the inter-disk space h, the size of the probe volume (0.81 mm in the axial direction) is not small compared to the boundary layer thicknesses and to the parameter h. Pressure is measured using 6 piezoresistive transducers. These transducers are highly accurate (0.05% in the range 10 to 40 C ) and combine both pressure sensors and temperature electronic compensations. They are xed to the stator at the following radial positions 0.093, 0.11, 0.14, 0.17, 0.2 and 0.23 m located on two rows because of geometrical constraints. Previous pressure measurements by embedded pressure gauges [36] showed that the pressure on the rotor and the one on the stator at the same radius are in fact identical within 2.5% accuracy. This is in fact a direct consequence of the TaylorProudman theorem which forbids axial gradients in rapidly rotating ows.

7
III. A. STATISTICAL MODELING

The dierential Reynolds Stress Model (RSM)

The ow studied here presents several complexities (high rotation rate, imposed throughow, wall eects, transition zones), which are severe conditions for turbulence modeling methods [3739]. Our approach is based on one-point statistical modeling using a low Reynolds number second-order full stress transport closure derived from the Launder and Tselepidakis [26] model and sensitized to rotation eects [30, 40]. Previous works [29, 30, 41] have shown that this level of closure was adequate in such ow congurations, while the usual k model, which is blind to any rotation eect presents serious deciencies. This approach allows for a detailed description of near-wall turbulence and is free from any eddy viscosity hypothesis. The general equation for the Reynolds stress tensor Rij can be written: dRij = Pij + Dij + ij dt where Pij , Dij , ij ,
ij , ij

+ Tij

(1)

and Tij respectively denote the production, diusion, pressure-strain

correlation, dissipation and extra terms.


T The diusion term Dij is split into two parts: a turbulent diusion Dij , which is inter-

preted as the diusion due to both velocity and pressure uctuations [42] and a viscous
diusion Dij , which can not be neglected in the low-Reynolds number region:

k T Dij = (0.22 Rkl Rij,l ),k Dij = Rij,kk

(2) (3)

In a classical way, the pressure-strain correlation term ij can be decomposed as below: ij = ij + ij + ij


(1) (1) (2) (w)

(4)

ij is interpreted as a slow nonlinear return to isotropy and is modeled as a quadratic development in the stress anisotropy tensor, with coecients sensitized to the invariants of anisotropy. This term is damped near the wall: 1 (1) ij = (c 1 aij + c1 (aik akj A2 ij )) 3 (5)

8 where aij denotes the stress anisotropy tensor and c 1 and c1 are two functions deduced from Crafts high-Reynolds number proposals [43] adapted for conned ows:

aij =

Rij 2 ij k 3
Re2 t 40

(6) ) (7) (8)

c 1 = (3.1 AA2 + 1)(1 e

c1 = 3.72 AA2 (1 e

Re2 40t

A = 1 9/8(A2 A3 ) is Lumleys atness parameter with A2 and A3 the second and third invariants of the anisotropy tensor. Ret = k 2 /() is the turbulent Reynolds number. The linear rapid part ij includes cubic terms. It can be written as:
(2)

1 Pkk (2) ij = 0.6(Pij Pkk ij ) + 0.3aij 3 Rkj Rli Rlk 0.2[ (Vk,l + Vl,k ) (Rik (Vj,l + jml m ) k k + Rjk (Vi,l + iml m ))] min(0.6, A)(A2 (Pij Dij ) + 3ami anj (Pmn Dmn )) (9)

with Pij = Rij Vj,k Rjk Vi,k and Dij = Rik Vk,j Rjk Vk,i . Since the slow part of the pressure-strain correlation is already damped near the wall, a wall correction ij is only applied to the rapid part. The form retained here is the one proposed by Gibson and Launder [44] with a strongly reduced numerical coecient. Moreover the classical length scale k 3/2 1 is replaced by k/(Rij ni nj )1/2 which is the length scale of the uctuations normal to the wall:
(w )

3 (2) (R) (w ) (2) (R) ij = 0.2[(km + km )nk nm ij (ik + ik )nk nj 2 k Rpq np nq 3 (2) (R) (kj + kj )ni nk ] 2 y

(10)

9 y is evaluated by the minimal distance of the current point from the four walls. The viscous dissipation tensor ij has been modeled in order to conform with the wall limits obtained from Taylor series expansions of the uctuating velocities [45]: ij = fA ij + (1 fA )(fs with fA , fs and ij dened as followed: Rij 2 + (1 fs )ij ) k 3 (11)

fA = e20A e
Re2 40t

Re2 t 20

(12) (13) (14)

fs = e (Rij + Rik nj nk + Rjk ni nk + Rkl nk nl ni nj ) ij = k 3 Rpq (1 + 2 np nq ) k

The extra term Tij accounts for implicit eects of the rotation on the turbulence eld, it contains additional contributions in the pressure-strain correlation, a spectral jamming term, inhomogeneous eects and inverse ux due to rotation, which impedes the energy cascade. This term allowed some improvements of results [30] in the Itoh et al. [16] calculation. Below is the proposal of Launder and Tselepidakis [26] for the dissipation rate equation : k d = c1 Rij Vi,j c2 f + (c Rij ,j + ,i ),i dt k k k + c3 Rjk Vi,jl Vi,kl + (c4 k,i ),i k
1/2 1/2

(15)

is the isotropic part of the dissipation rate = 2k,i k,i . c1 = 1, c2 = 1.92, c = 0.15, c3 = 2, c4 = 0.92 are four empirical constants and f is dened by: f = 1/(1 + 0.63 AA2 ). The kinetic turbulent energy equation is redundant in a RSM model but it is however still solved numerically in order to get faster convergence:

Tjj k dk = Rij Vi,j + + 0.22( Rij k,j + k,i ),i dt 2

(16)

It is veried after convergence that k is exactly 0.5Rjj .

10
B. Numerical method

The computational procedure is based on a nite volume method using staggered grids for mean velocity components with axisymmetry hypothesis in the mean. The computer code is steady elliptic and the numerical solution proceeds iteratively. A 140 80 mesh in the (r, z ) frame proved to be sucient in most of the cases considered in the present work to get grid-independent solutions [41]. However a more rened modeling 200 100 is necessary for higher rotation rates such as Re = 4.15 106 . About 20000 iterations (several hours on the NEC SX-5, IDRIS, Orsay, France) are necessary to obtain the numerical convergence of the calculation. In order to overcome stability problems, several stabilizing techniques have been introduced in the numerical procedure, such as those proposed by Huang and Leschziner [46]. Also, the stress component equations are solved using matrix block tridiagonal solution to enhance stability using non staggered grids.

C.

Boundary conditions

At the wall, all the variables are set to zero except for the tangential velocity V , which is set to r on rotating walls and zero on stationary walls. At the inlet, V is supposed to vary linearly from zero on the stationary wall up to r on the rotating wall. We recall that the inlet is close to the axis of the cavity when a centrifugal throughow is superimposed, whereas it is located at the periphery in the case of a centripetal throughow. When a throughow (centrifugal or centripetal) is enforced, a parabolic prole is then imposed for the axial velocity Vz at the inlet, with a given low level of turbulence intensity. In the outow section, the pressure is xed, whereas the derivatives for all the other independent quantities are set to zero if the uids leaves the cavity, and xed external values are imposed if the uid re-enters the cavity. In this case, the continuity equation is used to determine this inward or outward velocity component. The ow in the similarity area is practically not sensitive to the shape of proles of tangential and axial velocity components or to the intensity level imposed at the inlet. By multiplying by a factor 3 the turbulence intensity level imposed at the inlet, the change is about 0.08% on the maximum of the turbulent kinetic energy in the whole cavity. Moreover, these choices are justied by the wish to have a model as universal as possible.

11
IV. TURBULENT FLOW IN A CLOSED CAVITY

We study rst the turbulent ow in a closed rotor-stator system of aspect ratio G = 0.036 with no throughow. This basic ow belongs to the Batchelor type family: the two boundary layers are separated by a central rotating core. Two values of the rotational Reynolds number are here investigated: Re = 106 and Re = 4.15 106 . We dene the following dimensionless quantities: r = r/R2 , z = z/h, Vr = Vr /(r), V = V /(r) and Vz = Vz /(r). Note that z = 0 corresponds to the rotor side and z = 1 to the stator side.

A.

Structure of the mean ow

Figure 2 shows the structure of the mean ow in a closed cavity for Re = 106 . The ow is of Batchelor type. It is indeed clearly divided into three distinct zones: a centripetal boundary layer on the stator (B odewadt layer), a central rotating core and a centrifugal boundary layer on the rotor (Ekman or Von K arm an layer). In the B odewadt layer, the mean radial velocity is negative and the mean tangential velocity ranges between 0 and K r, with K the entrainment coecient of the uid. When one approaches the axis of the cavity, the B odewadt boundary layer thickness decreases and the minimum value of the radial velocity increases. The central core is characterized by a quasi zero radial velocity and a constant tangential velocity K r. K increases slightly from 0.38 at r = 0.44 to 0.45 at r = 0.8, whereas the size of the core decreases slightly with r . The Ekman layer is always centrifugal: the mean radial velocity is positive and the mean tangential velocity ranges between r and K r. When one approaches the axis of the cavity, the Ekman boundary layer thickness decreases and the maximum value of the radial velocity in this layer increases. The corresponding streamlines are displayed in gure 3a. Note that, for these three radial locations, there is an excellent agreement between the experimental data and the model predictions (g.2). These results are quite comparable with the predictions of Iacovides et al. [28] using a low-Re dierential second moment closure and the data of Cheah et al. [17] for Re = 1.6 106 and G = 0.127. Itoh et al. [16] found that K is sensitive to the radial location for Re = 106 and G = 0.08. It varies from 0.31 at r = 0.4 to 0.42 at r = 0.94. These authors notice that, for r 0.6, the Ekman boundary layer is not fully turbulent and the entrainment coecient K is then lower than 0.4.

12 By increasing the Reynolds number (g.3b, g.4), the structure of the mean ow does not change signicantly at the radial location r = 0.56. It always belongs to the Batchelor family with an entrainment coecient of the rotating core K also close to 0.45. Nevertheless, it can be noticed that the centrifugal force is weaker. Indeed, when the Reynolds number increases, the maximum value of the radial velocity in the Ekman layer decreases. The radial velocity prole is in very good agreement with the predictions of the algebraic stress model of Iacovides and Theofanopoulos [21] (Re = 4.4 106 , Cw = 0, G = 0.0255, r = 0.765): the maximum value of the radial velocity in the boundary layers are well predicted. They also bring out the existence of a rotating core with a quasi zero radial velocity. Nevertheless, the entrainment coecient determined by the authors K = 0.55 seems to be overestimated compared to the results of Poncet et al. [1].

B.

Pressure distributions

To complement the comparisons, we performed also pressure measurements by means of 6 pressure transducers. We choose to take as a reference the pressure measured at the outer radial position r = 0.92 and we dene the following pressure coecient: Cp =
2 P (r ) P (0.92). The dimensionless pressure is given by: P = 2P/(2 R2 ). In gure 5,

the pressure coecient is plotted versus the dimensionless radial position for Cw = 0 and two Reynolds numbers. As expected, the pressure decreases towards the center of the cavity: Cp is then always negative. At a given radius, it can be observed that Cp is almost the same in this range of Re. The case Re = 106 seems to show a greater discrepancy (g.5). However the normalization used in Cp magnies this apparent deviation. Indeed, the physical value of the pressure dierence between measurement and calculation is very small and remains within experimental accuracy. In a closed cavity, the radial velocity in the core is almost zero (g.2). For all the cases considered here, the axial velocity is almost zero too (see gure 4), that means that the mean ow is quasi two dimensional. So the Navier-Stokes equation for the tangential component reduces to the balance of the centrifugal force and of the radial pressure gradient: V2 /r = P/r. Using dimensionless quantities, the resulting equation linking the pressure coecient and the entrainment coecient is: dCp (r )/dr = 2K 2 r (17)

13 K being constant for both Reynolds numbers (g.2,4), the radial distributions of the pressure coecient are logically rather the same. This relation is a useful check to compare velocity and pressure measurements.

C.

Turbulence statistics

In this section, comparisons between the model results (RSM) and experimental data are given for three components of the Reynolds stress tensor. Poncet et al. [29] have already showed that the RSM model is the adequate level of closure to describe such complex ows, compared to a classical k model, which is blind to any rotation eect. Indeed it overestimates the turbulent intensities and fails to mimic the right proles. Comparisons for three components of the Reynolds stress tensor are given in gure 6 for Re = 106 at three radial locations. The RSM model provides good results compared to the experimental data even in the boundary layers but some discrepancies occur mainly in the low turbulence intensity regions where relaminarization is expected. In general, the turbulent intensities are rather weak in the considered case in the whole cavity. The turbulent intensity is mostly concentrated in the boundary layers. The Ekman layer is besides more turbulent than the B odewadt layer. In the rotating core, the Reynolds
stresses are weak. R12 is indeed almost zero, that means that there is no turbulent shear

stress in that zone. In the same way, there is no molecular viscous shear stress as the axial gradient of the radial velocity is zero in the core (g.2). Close to the axis (r = 0.44), we can notice a laminarization of the ow, which becomes locally turbulent, when one approaches the periphery of the cavity. Figure 7 gives an overview of the turbulent eld for Re = 4.15 106 at r = 0.56. By increasing the Reynolds number (g.6b,7), the turbulent intensities increase outside the
= vr vz /(r)2 , , Rrz Ekman layer. The three Reynolds shear stress tensor components Rr = v vz /(r)2 are negligible in the core. It conrms the existence of the inviscid rotating Rz

core. It means also that the production is almost zero in that area and that turbulence is
= vz2 /(r)2 component is rather weak too. only due to the diusion phenomenon. The Rzz

The model predictions are there in excellent agreement with the velocity measurements.

14
V. TURBULENT FLOW WITH CENTRIPETAL THROUGHFLOW

In this section, we study the turbulent ow inside a rotor-stator cavity of aspect ratio G = 0.036 when a centripetal throughow is superimposed on the basic ow. Three values of the ow rate coecient are studied: Cw = 1976, 5929, 9881 as well as two values of the rotational Reynolds number: Re = 106 and Re = 4.15 106 .

A.

Structure of the mean ow

Figure 8 shows the mean velocity proles for Re = 106 and Cw = 5929 at three radial locations. When an inward throughow is superimposed, the structure of the mean ow at the periphery preserves the properties of the ow in a closed cavity [13]. At r = 0.8, the ow is indeed of Batchelor type with two boundary layers separated by a central rotating core. When one approaches the axis of the cavity, the Ekman layer, which was centrifugal, becomes centripetal and the core rotates faster than the rotating disk. The ow is then centripetal whatever the axial location. K increases close to the axis because of the conservation of the angular momentum. The limit case is obtained for r = 0.56. A stagnation line is created on the rotor: the Ekman layer disappears and the core rotates at the same velocity as the rotor (g.10b). It is similar to that observed by Dijkstra and Van Heijst [47] and by Iacovides and Theofanopoulos [21]. When a radial centripetal throughow Cw = 3795 is superimposed of the turbulent basic ow Re = 6.9 105 , G = 0.0685, these last authors predict a limit case at r = 0.47. At a given radial location r = 0.56 and for Re = 106 , gure 9 shows the inuence of an increasing centripetal throughow on the structure of the mean ow. For Cw = 1976, the ow is of Batchelor type. The Ekman layer is centrifugal and the core rotates slower than the rotor. By increasing the inward throughow Cw = 9881, the Ekman layer becomes centripetal and the core rotates then faster than the rotating disk. The mean radial velocity is negative almost everywhere. The agreement between the experimental data and the model results is very satisfactory except for the prediction of K , which is slightly underestimated by the RSM model. When the value of the rotation rate increases from Re = 106 (g.9b, Cw = 5929, r = 0.56) to 4.15 106 (g.11), the entrainment coecient K decreases from 1 to 0.64. Figure 11

15 shows the mean velocity proles for Re = 4.15 106 and Cw = 5929 at r = 0.56. The ow is of Batchelor type: the Ekman becomes gets again centrifugal and the core rotates slower than the rotor. The axial velocity component is almost zero. The corresponding streamlines are displayed in gure 10d.

B.

The entrainment coecient K

The predictions of the RSM model have been also validated on experimental data measured using the two-component LDA. Poncet et al. [1] have shown analytically and experimentally that the entrainment coecient K of the rotating uid can be correlated, in the case of a centripetal throughow, to a local ow rate coecient: Cqr = according to a 5/7 power law: K = 2(a Cqr + b)5/7 1 (18)
2 Q ( r )1/5 2r3

with a and b experimental constants. In gure 12, several points deduced from the modeling results are plotted against the mean experimental K-curve. The constants deduced from the model a = 5.3 and b = 0.63 are very close to those obtained by the measurements a = 5.9 and b = 0.63. The maximum discrepancy in K-values between experiments and the numerical predictions is quite weak (less than 9%) and remains within the uncertainty range of both experimental and numerical approaches. Many subtle inuences coming from uncertainties in prerotation level or boundary layer prediction may contribute to explain this small discrepancy. This law is not sensitive to variations of the aspect ratio G, as far as the ows remain in the regime IV dened by Daily and Nece [11] (turbulent with separated boundary layers). These two coecients depend strongly on the prerotation level. Debuchy [24] studied indeed the case of a weak prerotation. These velocity measurements can be tted by the relation (18) with a = 2.8 and b = 0.46. For high values of Cqr , the core rotates faster than the rotating disk (K > 1). By increasing the throughow or decreasing the angular velocity of the rotor or by approaching the center of the cavity, the value of K increases: it could be due to the vortex stretching phenomenon. In our experiments, the core can rotate two times faster than the rotating disk. Debuchy et al. [18] have studied the turbulent ow (Re = 1.47 106 ) in a rotor-stator

16 cavity of large aspect ratio (G = 0.08) when a weak inward throughow (Cw = 188) is superimposed. The entrainment coecient is then very sensitive to the shrouding. For a small clearance, K is close to 0.38 at the periphery and depends slightly on Cw . On the contrary, at r = 0.53, K depends strongly on Cw and reaches 0.9 for Cw = 188.

C.

Pressure distributions

In gure 13, the pressure coecient Cp is plotted versus the dimensionless radial location for three centripetal throughows. As expected, the pressure decreases towards the center of the cavity: Cp is then always negative. At a given radius and for a given Reynolds number Re, it can be observed that Cp increases for increasing values of the ow rate Cw (in absolute value). That is in agreement with the measurements and the theoretical model of Debuchy et al. [18]. According to relation (17), we can determine the entrainment coecient from the value of Cp . We rst perform a polynomial t of the curve Cp versus r and then, we calculate by nite dierence the derivative of Cp to obtain K . Figure 14 compares the radial variations of K for the data series obtained by pressure and velocity measurements in the cases of centripetal throughows. As it can be seen, the results are in excellent agreement. The small dierences come in fact from the calculation of the derivative of Cp and from the hypothesis of zero radial velocity, which is less relevent for strong centripetal throughows. The pressure coecient Cp enables then to know the axial thrusts on the rotor.

D.

Turbulence statistics

Figure 15 shows the axial proles of three components of the Reynolds stress tensor for Re = 106 and Cw = 5929 at three radial locations. The model results are in excellent agreement with the experimental data. The turbulent intensities are well predicted apart
component, which is underestimated in the B odewadt layer. For a given from the Rr

Reynolds number and a given ow rate coecient, the ow is more turbulent close to the axis of rotation than at the periphery of the cavity. The B odewadt layer is besides more turbulent than the Ekman layer. The viscous shear stress is quasi zero in this layer, whereas it is large on the stator. The value of the normal Reynolds stress tensor components are

17 quite comparable in the whole cavity. The eect of the ow rate coecient on the turbulent eld is displayed in gure 16. A centripetal throughow increases the turbulent intensities. The turbulent levels are higher for a strong inward throughow (Cw = 9881) than for a weak throughow outside the Ekman layer. Indeed, the maximum values of the Reynolds stresses in the B odewadt layer and the turbulent levels in the core increase with increasing values of Cw . As expected, by increasing the Reynolds number from Re = 106 to 4.15 106 , the turbulent intensities increase noticeably (g.17). The shape of the turbulent proles does not change: the B odewadt layer is still more turbulent than the Ekman layer. But the values of the three Reynolds stresses have increased by a factor two.

VI.

TURBULENT FLOW WITH CENTRIFUGAL THROUGHFLOW

In this section, we study the turbulent ow inside a rotor-stator cavity of aspect ratio G = 0.036 when a centrifugal throughow is superimposed on the basic ow. Three values of the ow rate coecient are studied: Cw = 1976, 5929, 9881 as well as two values of the rotational Reynolds number: Re = 106 and Re = 4.15 106 .

A.

Structure of the ow

Figure 18 shows the inuence of the radial location on the mean ow when a centrifugal throughow is superimposed for Re = 106 and Cw = 5929. At the periphery r = 0.8, the ow preserves the properties of the ow without throughow: two boundary layers separated by a rotating core but, in this case, the core rotates slower than the rotor and than the case in the closed cavity (Cw = 0). The Ekman layer is centrifugal (Vr > 0) whereas the B odewadt layer is centripetal (Vr < 0). By approaching the center of the cavity, the core disappears and the B odewadt layer gets centrifugal. The ow is then fully centrifugal (V 0 and Vr > 0 everywhere) whatever the axial location z : it is a Stewartson type of ow. We can consider it as the connection of two ows over a single disk: a B odewadt type of ow (xed disk) and a Von K arm an type of ow (rotating disk). The Stewartson ow structure is composed of a single boundary layer on the rotor and a quasi zero tangential velocity outside (the corresponding streamline patterns are shown in gure 20b). This ow structure has been

18 observed by Chew [19] using a k turbulence model for Re = 3.4 106 , Cw = 5.4 104 at r = 1. In the same manner, when a weak centrifugal throughow is imposed (g.19a, g.20a), the ow keeps the same characteristics as in a closed cavity: two boundary layers separated by a central core, that is known as a Batchelor type of ow. The entrainment coecient K decreases. Note that the Batchelor type prole can appear only if the B odewadt layer is centripetal. By increasing the ow rate coecient Cw (g.19b, g.20b), the central core disappears and the ow gets centrifugal everywhere (Vr > 0). The proles for the tangential velocity are then of Stewartson type. For stronger throughows (g.19c), the radial velocity proles get closer to a channel ow like prole. Daily et al. [14] have observed two noticeable eects of a centrifugal throughow (Cw = 3510) on the turbulent ow (Re = 6.9 105 ) in a rotor-stator cavity of large aspect ratio (G = 0.069): the reduction in core rotation and the Ekman boundary layer growth with decreasing radius. The eect of the Reynolds number Re is shown in gure 21 compared to gure 18. By increasing Re, the B odewadt layer becomes centripetal and the central core reappears (K 0.2). It slows down the transition between a Batchelor and a Stewartson ow structure The transition between the Batchelor and Stewartson ow structures is observed either when one approaches the center of the cavity (r decreases) or by decreasing the ow rate coecient Cw (by increasing the centrifugal throughow) or by decreasing the Reynolds number. This transition is mainly due to the radial velocity, which is positive whatever the axial location for a Stewartson type of ow and, which vanishes for an axial location for a Batchelor type of ow. Note that all the model results are in excellent agreement with the experimental data in the case of a centrifugal throughow. as when r decreases (g.19a, g.21). The axial velocity component is almost zero.

B.

Transition diagram between a Batchelor and a Stewartson type of ow

In the present work, the Stewartson type of ow has never been observed in the closed cavity nor in the case of a centripetal throughow. On the contrary, when a centrifugal throughow is superimposed, the transition between the Batchelor and the Stewartson ow
2 e) structure can be characterized by considering a Rossby number dened as Ro = Q/(2R2

versus the radial location r (g.22) and for a given aspect ratio G = 0.036. Ro depends

19 on r according to a third degree polynomial t, which has been determined empirically and, which collapses all the experimental and numerical data in a single correlation law: Ro = 0.0088 0.0998r + 0.3048r2 0.4646r3 . The whole experimental data (centripetal and centrifugal cases) can be merged into a single curve giving the variations of K versus Cqr (g.23). The law (18) has been validated for some centripetal throughows (Batchelor type of ow) and four aspect ratios G. When a centrifugal throughow is superimposed, both type of ow structures are observed. As long as the ow is of Batchelor type, the equation (18) law is still valid (low absolute value of Cqr ). We can then calculate the radial pressure gradient and also know the axial thrust on the rotor. For stronger centrifugal throughows (higher negative values of Cqr ), the symmetry of the Batchelor type of ow is broken and the ow is then of Stewartson type. It is enclosed in the Ekman boundary layer. The variation law of K changes but the transition between these two types of ow is carried out in a continuous manner. Note that there is no core in the Stewartson type of ow and it is dicult to dene an entrainment coecient. So we can take K as the mean dimensionless tangential velocity outside the Ekman layer. Finally, for strong centrifugal throughow, K goes to an asymptotical value close to 0. Nguyen et al. [48] have shown numerically that a cavity with a large aspect ratio furthers a Stewartson type of ow, whereas a cavity with a small aspect ratio furthers a Batchelor ow structure. Our experiments and the ones of Poncet et al. [49] for G = 0.012 and G = 0.036 do not have brought out any dierence in the ow structure and the equation (18) law is still valid for these two aspect ratios. G does not seem to be a relevant parameter in the present study and for this range of aspect ratios.

C.

Pressure distributions

In the case of a centrifugal ow with Re = 106 , the pressure coecient is very close to zero. In gure 24, the pressure coecient is plotted versus the dimensionless radial location for Re = 4.15 106 . As expected, the pressure decreases towards the center of the cavity. At a given radius and for this Reynolds number, it can be observed that Cp decreases for increasing values of the ow rate, which is contrary to the case with a centripetal throughow. When Cw = 5929, the radial pressure gradient is close to zero for r lower than 0.56. The ow is then of Stewartson type and for greater r , the ow is of Batchelor type. That is

20 conrmed by the velocity proles (g.21).

D.

Turbulence statistics

Figure 25 exhibits the axial proles of three Reynolds stresses for Re = 106 and Cw = 5929 at three radial locations. As for the case with centripetal throughows, the turbulent
intensities increase from the periphery to the center of the cavity. At r = 0.44, Rrr and R

are almost constant in the central core, whereas their proles are nearly linear for greater r . Contrary to the case with centripetal throughows, the maximum of the turbulent intensities is conned in the Ekman layer and they vanish in the B odewadt layer. The inuence of a centrifugal throughow on the turbulent eld is quite negligible for
1/2 this range of ow rate coecients (g.26). Rrr and R 1/2

are respectively close to 0.06

and 0.07 in the Ekman layer. These two Reynolds stresses are quite comparable for these
centrifugal throughows. Rr is maximum in the rotating disk boundary layer and reaches

0.002 whatever the value of Cw . When an intermediate centrifugal throughow is superimposed Cw = 5929, the axial proles of the three Reynolds stresses are almost the same for Re = 106 and Re = 4.15 106 (g.25b, g.27) in the middle of the cavity r = 0.56. To conclude, the turbulent intensities in a rotor-stator cavity with outward throughow, depend essentially on the radial location. For all the considered cases, the RSM predictions are in very good agreement with the experimental data, even in the boundary layers. Figure 28 sums up the inuence of the throughow on the turbulent eld considering the repartition of the turbulent Reynolds number Ret = k 2 /( ) in the whole cavity for G = 0.036, Re = 106 and three values of the owrate coecients Cw . In a closed cavity (g.28b), a laminar region subsists (r 0.4). It appears also that the B odewadt layer becomes turbulent closer to the axis than the Ekman layer. That conrms the previous results of Cheah et al. [17]. Because of geometrical constraints, we can not perform LDA measurements in the low turbulence intensity region where relaminarization is expected. Nevertheless, gure 6a suggests that the radial location r = 0.44 is possibly the beginning of this laminar area. The present low Reynolds number closure gives an account of this zone. But, when a centrifugal (g.28a) or centripetal (g.28c) throughow is superimposed, strong gradients due to this throughow prevent the subsistence of a laminar area close to

21 the axis of the cavity.

VII.

CONCLUSION

New experimental investigations and extensive measurements have been performed and compared to numerical predictions to describe the ow in a rotor-stator system with or without throughow according to a large range of the ow control parameters: a rotational Reynolds number Re, a ow rate coecient Cw and the aspect ratio of the cavity G. In a closed cavity as well as in the case of a centripetal throughow, the ow belongs to the Batchelor family. It is divided into three distincts zones: two boundary layers separated by a central rotating core. The dimensionless tangential velocity in the core K depends on a local ow rate coecient (function of Re and Cw ) according to a useful general 5/7 power law determined by Poncet et al. [1]. When a centrifugal throughow is superimposed, Batchelor and Stewartson ow structures can be found. A Stewartson type of ow is conned in the rotating disk boundary layer. The tangential velocity is almost zero outside and the radial velocity is positive everywhere. The results for the mean ow can be synthetized by the hodograph on gure 29, which shows, for Re = 4.15 106 and three ow rate coecients, the Ekman spirals describing the changes in direction of the velocity vector through the cavity at a given radial location r = 0.56. The tangential velocity in the central core is attained when the radial velocity is approximately zero, where a concentration of points occurs. The proles falls between the typical fully turbulent behavior and the laminar Von K arm an solution presented by [34]. Moreover it conrms that the Stewartson ow is conned in the Ekman layer whereas the Batchelor ow structure is more symmetric. We succeeded in characterizing the transition at a given radial location between these two ow structures according to a Rossby number. We showed that it does not depend on the aspect ratio G. This continuous transition occurs as soon as the radial velocity is negative for at least one axial location. The study includes also turbulence measurements, which were seldom
and possible in previous works of the literature. It appears that the turbulent intensities Rrr in the B odewadt layer increase from the periphery to the center of the cavity when a Rr

throughow is superimposed. On the contrary, in a closed cavity, these turbulent intensities increase by increasing r (g.30). The predictions of the present RSM turbulence model are found here in excellent agreement with the velocity and pressure measurements over a

22 wide range of parameters including strong outow and inow. The RSM model appears as a valuable tool for describing the mean and turbulent elds of such complex ows.

VIII.

ACKNOWLEDGEMENT

Numerical computations were carried out on the NEC SX-5 (IDRIS, Orsay, France). Financial supports for the experimental approach from SNECMA Moteurs, Large Liquid Propulsion (Vernon, France) are also gratefully acknowledged.

[1] S. Poncet, M.P. Chauve, and P. Le Gal. Turbulent rotating disk with inward throughow. J. Fluid. Mech., 522:253262, 2005. [2] V.W. Ekman. On the inuence of the Earths rotation on ocean-currents. Arkiv. Mat. Astr. Fys., 2(11):152, 1905. [3] T. Von K arm an. Uber laminare und turbulente Reibung. Z. Angew. Math. Mech., 1:233252, 1921. odewadt. Die Drehstr omung u ber festem Grunde. Z. Angew. Math. Mech., 20:241253, [4] U.T. B 1940. [5] G.K. Batchelor. Note on a class of solutions of the Navier-Stokes equations representing steady rotationally-symmetric ow. Quat. J. Mech. and Appl. Math., 4(1):2941, 1951. [6] K. Stewartson. On the ow between two rotating coaxial disks. Proc. Camb. Phil. Soc., 49:333341, 1953. [7] H.O. Kreiss and S.V. Parter. On the swirling ow between rotating coaxial disks: existence and uniqueness. Commun. Pure Appl.Math., 36:5584, 1983. [8] G.L. Mellor, P.J. Chapple, and V.K. Stokes. On the ow between a rotating and a stationary disk. J. Fluid. Mech., 31(1):95112, 1968. [9] P.J. Zandbergen and D. Dijkstra. Von K arm an swirling ows. Ann. Rev. Fluid Mech., 19:465 491, 1987. [10] J.F. Brady and L. Durlofsky. On rotating disk ow. J. Fluid. Mech., 175:363394, 1987. [11] J.W. Daily and R.E. Nece. Chamber dimension eects on induced ow and frictional resistance of enclosed rotating disks. ASME J. Basic Eng., 82:217232, 1960.

23
[12] G.N. Lance and M.H. Rogers. The axially symetric ow of a viscous uid between two innite rotating disks. Proc. R. Soc. London A, 266:109121, 1962. [13] J.M. Owen and R.H. Rogers. Flow and Heat Transfer in Rotating-Disc Systems - Vol.1: Rotor-Stator Systems. Ed. Morris, W.D. John Wiley and Sons Inc., New-York, 1989. [14] J.W. Daily, W.D. Ernst, and V.V. Asbedian. Enclosed rotating disks with superposed throughow. Technical Report 64, M.I.T, department of civil engineering, 1964. [15] M. Firouzian, J.M. Owen, Pincombe J.R., and R.H. Rogers. Flow and heat transfer in a rotating cylindrical cavity with a radial inow of uid. part 2: Velocity, pressure and heat transfer measurements. Int. J. Heat Fluid Flow, 7(1):2127, 1986. [16] M. Itoh, Y. Yamada, S. Imao, and M. Gonda. Experiments on turbulent ow due to an enclosed rotating disk. pages 659668, New-York, 1990. Proc. Int. Symp. on Engineering Turbulence Modeling and Experiments, Ed. W. Rodi and E.N. Galic, Elsevier. [17] S.C. Cheah, H. Iacovides, D.C. Jackson, H. Ji, and B.E. Launder. Experimental investigation of enclosed rotor-stator disk ows. Exp. Therm. Fluid Sci., 9:445455, 1994. [18] R. Debuchy, A. Dyment, H. Muhe, and P. Micheau. Radial inow between a rotating and a stationary disc. Eur. J. Mech. B/Fluids, 17(6):791810, 1998. [19] J.W. Chew. Prediction of ow in rotating disc systems using the k turbulence model. In ASME 84-GT-229, Amsterdam, 1984. Gas Turbine Conference. [20] J.W. Chew and C.M. Vaughan. Numerical predictions of ow induced by an enclosed rotating disk. In ASME 88-GT-127, Amsterdam, 1988. 33rd Gas Turbine and Aeroengine Congress. [21] H. Iacovides and I.P. Theofanopoulos. Turbulence modeling of axisymmetric ow inside rotating cavities. Int. J. Heat Fluid Flow, 12(1):211, 1991. [22] H. Iacovides and P. Toumpanakis. Turbulence modelling of ow in axisymmetric rotor-stator systems. In 5th Int. Symp. on Rened Flow Modelling and Turbulence Measurements, Presses de lEcole Nationale des Ponts et Chauss ees, 1993. [23] R. Schiestel, L. Elena, and T. Rezoug. Numerical modeling of turbulent ow and heat transfer in rotating cavities. Numer. Heat Transfer A, 24(1):4565, 1993. [24] R. Debuchy. Ecoulement turbulent avec aspiration radiale entre un disque xe et un disque tournant. PhD thesis, Universit e des Sciences et Technologies de Lille, 1993. [25] L. Elena and R. Schiestel. Turbulence modeling of conned ow in rotating disk systems. AIAA J., 33(5):812821, 1995.

24
[26] B.E. Launder and D.P. Tselepidakis. Application of a new second-moment closure to turbulent channel ow rotating in orthogonal mode. Int. J. Heat Fluid Flow, 15(1):210, 1994. [27] K. Hanjalic and B.E. Launder. Contribution towards a Reynolds-stress closure for low-

Reynolds number turbulence. J. Fluid. Mech., 74(4):593610, 1976. [28] H. Iacovides, K.S. Nikas, and M.A.F. Te Braak. Turbulent ow computations in rotating cavities using low-Reynolds-number models. In ASME 96-GT-159, Birmingham, UK, 1996. Int. Gas Turbine and Aeroengine Congress & Exhibition. [29] S. Poncet, R. Schiestel, and M.-P. Chauve. Turbulence modelling and measurements in a rotor-stator system with throughow. Sardinia, Italy, 2005. accepted for publication in the proceedings of the ERCOFTAC International Symposium on Engineering Turbulence Modelling and Measurements. [30] L. Elena and R. Schiestel. Turbulence modeling of rotating conned ows. Int. J. Heat Fluid Flow, 17:283289, 1996. [31] J. Fromm. Understanding turbulence through Navier-Stokes computation of ow between rotating disks. Technical report, IBM Almaden Research Center, 1987. IBM Research Report 5999 RJ Mathematics/Physics. er e, and O. Daube. Axisymmetric numerical simulations of turbulent [32] R. Jacques, P. Le Qu ow in a rotor stator enclosures. Int. J. Heat Fluid Flow, 23:381397, 2002. [33] E. Serre, P. Bontoux, and B.E. Launder. Direct numerical simulation of transitional turbulent ow in a closed rotor-stator cavity. Flow, Turbulence and Combustion, 69:3550, 2002. [34] M. Lygren and H.I. Andersson. Turbulent ow between a rotating and a stationary disk. J. Fluid. Mech., 426:297326, 2001. [35] S. Poncet, M.-P. Chauve, and P. Le Gal. Study of the entrainment coecient of the uid in a rotor-stator cavity. Lille, France, 2005. Proceedings of the European Turbomachinery Conference. [36] R.M. Gassiat. Etude exp erimentale d ecoulements centrip` etes avec pr erotation dun uide conn e entre un disque tournant et un carter xe. PhD thesis, Universit e de la M editerran ee Aix-Marseille II, 2000. [37] P. Bradshaw. The analogy between the streamline curvature and buoyancy in turbulent shear ow. J. Fluid. Mech., 36:177191, 1969. [38] C. Cambon and L. Jacquin. Spectral approach to non isotropic turbulence subjected to

25
rotation. J. Fluid. Mech., 202:295317, 1989. [39] L. Jacquin, O. Leuchter, C. Cambon, and J. Mathieu. Homogeneous turbulence in the presence of rotation. J. Fluid. Mech., 220:152, 1990. [40] R. Schiestel and L. Elena. Modeling of anisotropic turbulence in rapid rotation. Aerospace Science and Technology, 7:441451, 1997. [41] L. Elena. Mod elisation de la turbulence inhomog` ene en pr esence de rotation. PhD thesis, Universit e Aix-Marseille I-II, 1994. [42] B.J. Daly and F.H. Harlow. Transport equation for turbulence. Phys. Fluids A, 13(11):2634 2649, 1970. [43] T.J. Craft. Second-moment modelling of turbulent scalar transport. PhD thesis, University of Manchester, 1991. [44] M. Gibson and B.E. Launder. Ground eects on pressure uctuations in the atmospheric boundary layer. J. Fluid. Mech., 86(3):491511, 1978. [45] B.E. Launder and W.C. Reynolds. Asymptotic near-wall stress dissipation rates in a turbulent ow. Phys. Fluids A, 26(5):11571158, 1983. [46] P.G. Huang and M.A. Leschziner. Stabilization of recirculating ow computations performed with second moments closures and third order discretization. Cornell University, Ithaca, NY, 1985. Vth Int. Symp. on Turbulent Shear Flow. [47] D. Dijkstra and G.J.F. Van Heijst. The ow between two nite rotating disks enclosed by a cylinder. J. Fluid. Mech., 128:123154, 1983. [48] N.D. Nguyen, J.P. Ribault, and P. Florent. Multiple solutions for ow between coaxial disks. J. Fluid. Mech., 68(2):369388, 1975. [49] S. Poncet, R. Schiestel, and M.P. Chauve. Centrifugal ow in a rotor-stator cavity. accepted for publication in J. Fluid Eng., 2005.

26 Fig.1: Schematic representation of the experimental set-up and notations: R1 = 38, R2 = 250, R3 = 253, R4 = 55 and 3 h 12 mm. Fig.2: Mean velocity proles for Re = 106 and Cw = 0 at three radial locations: (a) r = 0.44, (b) r = 0.56, (c) r = 0.8; () RSM model, () experimental data.
2 Fig.3: Eect of the Reynolds number on the streamlines = /(R2 ) for Cw = 0

(RSM), 15 regularly spaced intervals: (a) Re = 106 , 0 0.017, (b) Re = 4.15 106 , 0 0.014. Fig.4: Mean velocity proles for r = 0.56, Cw = 0 and Re = 4.15 106 ; () RSM model, () experimental data. Fig.5: Radial pressure distributions for Cw = 0 and two Reynolds numbers: (,) Re = 106 and ( ,) Re = 4.15 106 (the symbols represent the experimental data and the lines the model results). Fig.6: Proles of the Reynolds stress tensor components for Re = 106 , and Cw = 0 at three radial locations: (a) r = 0.44, (b) r = 0.56, (c) r = 0.8; () RSM model, () experimental data. Fig.7: Axial proles of the six Reynolds stress tensor components for Cw = 0 and Re = 4.15 106 at r = 0.56; () RSM model, () experimental data. Fig.8: Mean velocity proles for Re = 106 , Cw = 5929 at three radial locations: (a) r = 0.44, (b) r = 0.56, (c) r = 0.8; () RSM model, () experimental data. Fig.9: Mean velocity proles for Re = 106 , r = 0.56 and three ow rate coecients: (a) Cw = 1976, (b) Cw = 5929, (c) Cw = 9881; () RSM model, () experimental data.
2 ) patterns for some centripetal throughow values Fig.10: Streamlines = /(R2

(RSM), 20 regularly spaced intervals: (a) Re = 106 , Cw = 1976, 0.009 0.016, (b) Re = 106 , Cw = 5929, 0.027 0.012, (c) Re = 106 , Cw = 9881, 0.045 0.007, (d) Re = 4.15 106 , Cw = 5929, 0.009 0.008. Fig.11: Mean velocity proles for r = 0.56, Cw = 5929 and Re = 4.15 106 ; () RSM model, () experimental data.

27 Fig.12: Mean K curve: () experimental data, () analytical law of Poncet et al. [1], ( ) model results, () numerical correlation curve K = 2 (5.3 Cqr + 0.63)5/7 1. Fig.13: Radial pressure distributions for Re = 106 and three centripetal throughows: () model results and experimental data for () Cw = 1976, ( ) Cw = 5929, ( ) Cw = 9881. Fig.14: Radial proles of K for Re = 106 and three centripetal throughows; Comparison between pressure and velocity measurements. Fig.15: Proles of the Reynolds stress tensor components for Re = 106 , Cw = 5929 at three radial locations: (a) r = 0.44, (b) r = 0.56, (c) r = 0.8; () RSM model, () experimental data. Fig.16: Eect of the ow rate on the axial proles of the Reynolds stress tensor components for Re = 106 at r = 0.56: (a) Cw = 1976, (b) Cw = 5929, (c) Cw = 9881; () RSM model, () experimental data. Fig.17: Axial proles of the Reynolds stress tensor components for Cw = 5929 and Re = 4.15 106 at r = 0.56; () RSM model, () experimental data. Fig.18: Mean velocity proles for Re = 106 and Cw = 5929 at three radial locations: (a) r = 0.44, (b) r = 0.56, (c) r = 0.8; () RSM model, () experimental data. Fig.19: Mean velocity proles for Re = 106 , r = 0.56 and three ow rate coecients: (a) Cw = 1976, (b) Cw = 5929, (c) Cw = 9881; () RSM model, () experimental data.
2 Fig.20: Eect of the ow rate on the streamlines = /(R2 ) for Re = 106 (RSM),

15 regularly spaced intervals: (a) Re = 106 , Cw = 1976, 0 0.024, (b) Re = 106 , Cw = 5929, 0 0.032, (c) Re = 106 , Cw = 9881, 0 0.046, (d) Re = 4.15 106 , Cw = 5929, 0.001 0.019. Fig.21: Mean velocity proles for Cw = 5929 and Re = 4.15 106 at r = 0.56; () RSM model, () experimental data. Fig.22: Transition diagram (r ,Ro) between Batchelor and Stewartson ow structure: () experimental data, ( ) RSM model.

28 Fig.23: Experimental mean K curve compared to the analytical law () of Poncet et al. [1] for four aspect ratios: () G = 0.012, (2) G = 0.024, () G = 0.036, ( ) G = 0.048. Fig.24: Radial pressure distributions for Re = 4.15 106 and three centrifugal throughows: () model results, (2) Cw = 1976, () Cw = 5929, ( ) Cw = 9881. Fig.25: Axial proles of three components of the Reynolds stress tensor for Re = 106 and Cw = 5929 at three radial locations: (a) r = 0.44, (b) r = 0.56, (c) r = 0.8; () RSM model, () experimental data. Fig.26: Eect of the ow rate on the axial proles of three components of the Reynolds stress tensor for Re = 106 at r = 0.56: (a) Cw = 1976, (b) Cw = 5929, (c) Cw = 9881; () RSM model, () experimental data. Fig.27: Axial proles of three components of the Reynolds stress tensor for Cw = 5929 and Re = 4.15 106 at r = 0.56; () RSM model, () experimental data. Fig.28: Eect of the ow rate on the iso-turbulent Reynolds number Ret = k 2 /( ) for Re = 106 (RSM), 20 regularly spaced intervals: (a) Cw = 5929, Ret 1942, (b) Cw = 0, Ret 352, (c) Cw = 5929, Ret 7206. Fig.29: Hodograph, r = 0.56, Re = 4.15 106 : (, ) Cw = 5929, (,) Cw = 0, (,2) Cw = 5929. Fig.30: Radial proles of the Reynolds stress tensor components in the B odewadt layer at z = 0.946 for Re = 106 : (.,2) Cw = 5929, (,) Cw = 0, (, ) Cw = 5929.

29

z
R4
STATOR Interdisk space ROTOR

Q r

breakthrough crown

TANK

R1

R2 R3

M
Figure 1: Poncet et al., Phys. Fluids.

30

(a)
1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0

(b)
1 0.8 0.6 0.4 0.2 0

(c)

0.5

0.5

0.5

V*
1 0.8 0.6 0.4 0.2 0 0.2 1 0.8 0.6 0.4 0.2 0 0.2

V*
1 0.8 0.6 0.4 0.2 0 0.2

V*

0.2

0.2

0.2

V* r

V* r
Figure 2: Poncet et al., Phys. Fluids.

V* r

31

1 0.8 0.6

(a)
0.4 0.2 0

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1 0.8 0.6

(b)
0.4 0.2 0

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Figure 3: Poncet et al., Phys. Fluids.

32
1 1 1

0.9

0.9

0.9

0.8

0.8

0.8

0.7

0.7

0.7

0.6

0.6

0.6

z* 0.5
0.4

0.5

0.5

0.4

0.4

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

0.1

0.5

0 0.1

0.1

0.2

3 x 10
3

V*

V* r
Figure 4: Poncet et al., Phys. Fluids.

V* z

33
0.05

0.05

0.1

Cp
0.15

0.2

0.25

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Figure 5: Poncet et al., Phys. Fluids.

34
1 1 1

(a) z

0.5

0.5

0.5

0 1

0.02

0.04

0.06

0 1

0.02

0.04

0.06

0 10 1

5 z*

0 x 10
4

* (b) z 0.5

0.5

0.5

0 1

0.02

0.04

0.06

0 1

0.02

0.04

0.06

0 10 1

0 x 10
4

* (c) z 0.5

0.5

0.5

0.02

0.04

0.06

0.02

0.04

0.06

0 10

0 x 10
4

R*1/2 rr

R*1/2
Figure 6: Poncet et al., Phys. Fluids.

R* r

35
1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0

0.02

0.04

0.06

0.02

0.04

0.06

0.02

0.04

0.06

R*1/2 rr
1 0.8 0.6 0.4 0.2 0 8 1 0.8 0.6 0.4 0.2 0 8

R*1/2
1 0.8 0.6 0.4 0.2 0 8

R*1/2 zz

0 x 10

2
4

0 x 10

2
4

0 x 10

2
4

R* r

R* rz

R* z

Figure 7: Poncet et al., Phys. Fluids.

36

(a)
1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0

(b)
1 0.8 0.6 0.4 0.2 0

(c)

V*
1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0

V*
1 0.8 0.6 0.4 0.2 0

V*

0.6 0.4 0.2

0.6 0.4 0.2

0.6 0.4 0.2

V* r

V* r
Figure 8: Poncet et al., Phys. Fluids.

V* r

37
(a)
1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0

(b)
1 0.8 0.6 0.4 0.2 0

(c)

z*

0.5

1.5

0.5

1.5

0.5

1.5

V*
1 0.8 0.6 1 0.8 0.6 0.4 0.2 0

V*
1 0.8 0.6 0.4 0.2 0

V*

z*
0.4 0.2 0

0.4

0.2

0.4

0.2

0.4

0.2

V* r

V* r
Figure 9: Poncet et al., Phys. Fluids.

V* r

38

(a) 0.5
0 1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(b) 0.5
0 1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(c) 0.5
0 1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

(d) 0.5
0 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 10: Poncet et al., Phys. Fluids.

39
1 1 1

0.9

0.9

0.9

0.8

0.8

0.8

0.7

0.7

0.7

0.6

0.6

0.6

z*

0.5

0.5

0.5

0.4

0.4

0.4

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

0.1

0.5

0 0.2

0.1

0.1

3 x 10
3

V*

V* r
Figure 11: Poncet et al., Phys. Fluids.

V* z

40

Figure 12: Poncet et al., Phys. Fluids.

41
0

0.2

0.4

Cp

0.6

0.8

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Figure 13: Poncet et al., Phys. Fluids.

42

Figure 14: Poncet et al., Phys. Fluids.

43
1 1 1

(a) z

0.5

0.5

0.5

0 1

0.05

0.1

0 1

0.05

0.1

0 1

4 z

2
*

0 x 10
3

(b) z

0.5

0.5

0.5

0 1

0.05

0.1

0 1

0.05

0.1

0 1

2 x 10

0
3

* (c) z

0.5

0.5

0.5

0.05

0.1

0.05

0.1

0 x 10
3

R*1/2 rr

R*1/2
Figure 15: Poncet et al., Phys. Fluids.

R* r

44
1 1 1

(a) z

0.5

0.5

0.5

0 1

0.05

0.1

0 1

0.05

0.1

0 4 1

0 x 10
3

(b) z

0.5

0.5

0.5

0 1

0.05

0.1

0 1

0.05

0.1

0 4 1

0 x 10
3

* (c) z 0.5

0.5

0.5

0.05

0.1

0.05

0.1

0 4

0 x 10
3

R*1/2 rr

R*1/2
Figure 16: Poncet et al., Phys. Fluids.

R* r

45
1 1 1

0.9

0.9

0.9

0.8

0.8

0.8

0.7

0.7

0.7

0.6

0.6

0.6

*
0.5 0.5 0.5

0.4

0.4

0.4

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

0.1

0.05

0.1

0.15

0.05

0.1

0.15

0 4

0 x 10
3

R*1/2 rr

R*1/2
Figure 17: Poncet et al., Phys. Fluids.

R* r

46

(a)
1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0

(b)
1 0.8 0.6 0.4 0.2 0

(c)

0.5

0.5

0.5

V*
1 0.8 0.6 0.4 0.2 0 0.1 1 0.8 0.6 0.4 0.2 0 0.1

V*
1 0.8 0.6 0.4 0.2 0 0.1

V*

0.1

0.2

0.1

0.2

0.1

0.2

V* r

V* r
Figure 18: Poncet et al., Phys. Fluids.

V* r

47
(a)
1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0

(b)
1 0.8 0.6 0.4 0.2 0

(c)

z*

0.5

0.5

0.5

V*
1 0.8 0.6 1 0.8 0.6 0.4 0.2 0 0.1

V*
1 0.8 0.6 0.4 0.2 0 0.1

V*

0.4 0.2 0 0.1

0.1

0.2

0.1

0.2

0.1

0.2

V* r

V* r
Figure 19: Poncet et al., Phys. Fluids.

V* r

48

(a)

0.5

0 1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

(b)

0.5

0 1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

(c)

0.5

0 1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

(d)

0.5

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Figure 20: Poncet et al., Phys. Fluids.

49
1 1 1

0.9

0.9

0.9

0.8

0.8

0.8

0.7

0.7

0.7

0.6

0.6

0.6

z*

0.5

0.5

0.5

0.4

0.4

0.4

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

0.1

0.5

0 0.1

0.1

0.2

3 x 10
3

V*

V* r
Figure 21: Poncet et al., Phys. Fluids.

V* z

50

Figure 22: Poncet et al., Phys. Fluids.

51

Figure 23: Poncet et al., Phys. Fluids.

52

0.02

0.04

Cp

0.06

0.08

0.1

0.12 0.3

0.4

0.5

0.6

0.7

0.8

0.9

r*
Figure 24: Poncet et al., Phys. Fluids.

53
1 1 1

* (a) z 0.5

0.5

0.5

0 1

0.02

0.04

0.06

0.08

0 1

0.02

0.04

0.06

0.08

0 3 1

1 z
*

0 x 10

1
3

* (b) z 0.5

0.5

0.5

0 1

0.02

0.04

0.06

0.08

0 1

0.02

0.04

0.06

0.08

0 3 1

0 x 10

1
3

* (c) z 0.5

0.5

0.5

0.02

0.04

0.06

0.08

0.02

0.04

0.06

0.08

0 3

0 x 10

1
3

R*1/2 rr

R*1/2
Figure 25: Poncet et al., Phys. Fluids.

R* r

54
1 1 1

* (a) z 0.5

0.5

0.5

0 1

0.05

0.1

0 1

0.05

0.1

0 3 1

0 x 10
3

* (b) z 0.5

0.5

0.5

0 1

0.05

0.1

0 1

0.05

0.1

0 3 1

0 x 10
3

* (c) z 0.5

0.5

0.5

0.05

0.1

0.05

0.1

0 3

0 x 10
3

R*1/2 rr

R*1/2
Figure 26: Poncet et al., Phys. Fluids.

R* r

55
1 1 1

0.9

0.9

0.9

0.8

0.8

0.8

0.7

0.7

0.7

0.6

0.6

0.6

z*

0.5

0.5

0.5

0.4

0.4

0.4

0.3

0.3

0.3

0.2

0.2

0.2

0.1

0.1

0.1

0.02

0.04

0.06

0.08

0.02

0.04

0.06

0.08

0 3

0 x 10
3

R*1/2 rr

R*1/2
Figure 27: Poncet et al., Phys. Fluids.

R* r

56

(a) 0.5
0 1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

(b) 0.5
0 1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

(c) 0.5
0

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

r*
Figure 28: Poncet et al., Phys. Fluids.

57

Figure 29: Poncet et al., Phys. Fluids.

58
0.2 0.15

*1/2 Rr r 0.1
0.05 0 0.2 0.15 0.3 0.4 0.5 0.6 0.7 0.8 0.9

R*1/2

0.1 0.05 0 0.3 0.4 0.5 0.6 0.7 0.8 0.9

R* 0.01 r
0.02 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Figure 30: Poncet et al., Phys. Fluids.

S-ar putea să vă placă și