Sie sind auf Seite 1von 29

Fluid Dynamics Research 38 (2006) 145 173

DNS and LES of some engineering ows


Wolfgang Rodi
Institute for Hydromechanics, University of Karlsruhe, D-76128 Karlsruhe, Germany Received 20 February 2004; received in revised form 12 August 2004; accepted 2 November 2004 Communicated by F. Hamba

Abstract In this paper, direct numerical simulations (DNS) and large eddy simulations (LES) of three engineering ows carried out in the authors research group are presented. The rst example, simulated both with DNS and LES, is the ow in a low-pressure turbine cascade with wakes passing periodically through the cascade channel. In this situation, the laminarturbulent transition of the boundary layers on the blade surfaces, which is strongly inuenced by the passing wakes, is of special interest. Next, LES of the ow past the Ahmed body is presented, which is a car model with slant back. In spite of the fairly simple geometry, the ow around the model has many features of the complex, fully 3D ow around real cars. The third example, for which LES is presented, is the ow past a surface mounted circular cylinder of height-to-diameter ratio of 2.5. In this case also complex 3D ow develops with interaction of various vortices behind the cylinder. By means of these examples, the paper shows that complex turbulent ows of engineering relevance can be predicted realistically by DNS and LES, albeit at large cost. The methods are particularly suited and superior to RANS methods for situations where unsteadiness like shedding and large-scale structures dominate the ow, and DNS has evolved into an important tool for studying transition mechanisms. 2005 The Japan Society of Fluid Mechanics and Elsevier B.V. All rights reserved.
Keywords: Numerical ow simulation; DNS; LES; Turbulence; Transition; Turbine cascades; Car model; Finite-height cylinder

1. Introduction Most engineering ows are turbulent, i.e. they carry irregular, uctuating turbulent motions which contribute signicantly to the transport of momentum, heat and mass. These motions, therefore, affect
E-mail address: rodi@uka.de. 0169-5983/$30.00 2005 The Japan Society of Fluid Mechanics and Elsevier B.V. All rights reserved. doi:10.1016/j.uiddyn.2004.11.003

146

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

greatly the distribution of velocity, temperature and concentration over the ow eld and consequently the forces exerted by the ow, the mixing, dilution and heat and mass transfer. Turbulence can have positive or negative effects, and it is important for engineers to be able to predict the effects in the design process. Hence, the simulation of turbulence plays an essential part in most CFD calculations. Unfortunately, the turbulent motions are very complex as they are always three-dimensional (3D), unsteady and fairly irregular. The motion consists of eddies with a wide spectrum of sizes, ranging from large eddies of the size of the ow domain corresponding to low-frequency uctuations to much smaller eddies, at which dissipation takes place, corresponding to high-frequency uctuations. Various methods are available to simulate the turbulent motions and their effect on the ow development, heat and mass transfer. Here, the three main types of methods are briey introduced. All the details of the complex turbulent uctuating motion are governed by the unsteady 3D NavierStokes equations together with the continuity equation and a method that solves numerically these equations without introducing any model is called direct numerical simulation (DNS). Since with this method all motions need to be resolved, the size of the numerical mesh must be smaller than the size of the small-scale motion where dissipation takes place. As the relation of this to the size of the large-scale motion and hence the ow domain varies inversely with the Reynolds number and the calculations have to be always 3D, the number of grid points and the cost required increase roughly with Re3 . Hence, DNS is possible only for ows with relatively low Reynolds numbers. In large eddy simulations (LES) only the larger-scale turbulent motions that can be resolved on a given grid are simulated explicitly by solving the 3D time-dependent NavierStokes equations; the small-scale motions that cannot be resolved need to be accounted for by a subgrid-scale model. LES can be used to calculate high Reynolds numbers ows, but in this case the near-wall regions cannot be properly resolved and a special near-wall treatment has to be introduced. In contrast to LES, where a model has to be introduced only for the small-scale motion, calculations employing the Reynolds averaged NavierStokes (RANS) equations need a model to account for the entire turbulent spectrum as all turbulent uctuations are averaged out from the equations solved. This approach requires much less computational effort and is hence the method used in most practical calculations today, but since all effects of turbulence have to be accounted for by a model, this is a difcult task in view of the complexity and rich variety of different turbulent ow phenomena occurring. Engineering ows are often rather complex, especially when they involve separation and reattachment of the ow, large-scale eddying turbulent structures that dominate the turbulent transport, unsteady processes prevailing such as vortex shedding or bimodal unsteady behaviour, and transition from laminar to turbulent ow which can happen in various ways, subject to a wide variety of possible disturbances. The simulation of such processes is difcult with RANS models and hence these models had limited success in such complex situations. For these, DNS and LES are principally better suited because they simulate directly the turbulent motions, all of them in DNS and the larger ones that contribute most to the transport of momentum, heat and mass in LES. DNS requires no model at all while LES only a model for the small-scale motion and hence carries much less burden of the overall simulation and has an easier task because this motion is less inuenced by the boundary conditions so that it is easier to derive a universal model. As was mentioned already, DNS is limited to relatively low Reynolds number ows and even then the computing resources required are enormous. Hence, this is not a method for everyday engineering calculations, but it is a very useful tool for studying the details of turbulence and especially the mechanism of laminar to turbulent transition and for generating data which can be used as basis for testing and

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

147

developing further engineering calculation models. Also, there is an area where ows of practically relevant Reynolds numbers can already be simulated by DNS, namely in turbomachinery cascades, as will be shown later in this paper. LES can also be used for higher Reynolds number ows, and in fact away from walls the length scale of the turbulent eddies containing most of the energy and contributing most to the momentum, heat and mass transfer that must be resolved is virtually independent of the Reynolds number, so that in this area LES does not have a Reynolds number problem. However, near walls the length scale decreases with increasing Re so that the number of grid points required to resolve the near-wall zone adequately increases with Re2 . Again such calculations are not possible at high Reynolds numbers so that in order to apply LES, some near-wall modelling approximating the overall effects of turbulence near the wall must be introduced. One possibility is to use wall functions bridging the viscous sublayer and another the use of RANS modelling in the near-wall zone as used in the detached eddy simulation (DES) approach of Spalart et al. (1997). Although LES is also computationally expensive since strictly always a 3D unsteady calculation has to be carried out, it nds its way already into commercial CFD codes and is used in some practical calculations, especially for situations where large-scale structures dominate the ow behaviour and where unsteadiness prevails. LES also plays an increasing role in the area of computational aero acoustics where it provides the noise sources for such calculations. In the following, a few examples of DNS and LES of engineering ows performed in the authors research group will be presented and discussed.

2. Methods used The ow examples in this paper were simulated by solving numerically the incompressible 3D NavierStokes equations together with the continuity equation. In the case of DNS, the original unsteady equations were solved without introducing any model. In the LES, the equations are for ltered quantities where the lter width is determined by the grid size so that the resolved quantities are basically averages over the control volumes formed by the numerical mesh. The ltering or averaging introduces subgrid-scale (SGS) stress terms which are modelled by relating them via a SGS eddy viscosity to the local gradients of the resolved velocities. In the application examples, the SGS eddy viscosity is either determined with the original Smagorinsky model relating it to the strain rate and the average mesh size using a constant coefcient. Alternatively, a dynamic version of this model is used in which the coefcient is determined from the information available from the smallest resolved scales (Germano et al., 1991). Therefore, with this SGS model, in LES basically the same equations are solved as in DNS. Only in the latter the molecular viscosity is used while in LES the sum of molecular and SGS eddy viscosity appears instead. In some of the LES reported below, a near-wall model was used which is described when the application example is introduced. The numerical solution of the governing equations was performed with the nite-volume code LESOCC (Breuer and Rodi, 1996) or with its fully parallelised successor LESOCC2 developed at the Institute for Hydromechanics. These codes solve the equations on body-tted curvilinear block-structured grids using second-order central differences for the discretisation of the convection and diffusion terms. Time advancement is accomplished by a three stage RungeKutta scheme. Conservation of mass is achieved by the SIMPLE algorithm, with the pressure correction equation being solved by the SIP procedure.

148

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

3. Flow past turbine blade with incoming wakes The rst example concerns the ow in a low pressure turbine cascade. Due to the rotor/stator interaction, the ow is unsteady with wakes generated by the preceding row of blades passing through the cascade. These wakes have a strong inuence on the ow and through their interaction with the blade boundary layer also on the development of the latter and especially the laminar turbulent transition. In low pressure turbines, the Reynolds number is relatively low so that considerable portions of the blade boundary layer are laminar or only intermittently turbulent when the wakes pass. This phenomenon is difcult to simulate by RANS methods and is more suited to be dealt with by DNS and LES. Indeed, DNS can be used to study the ow for this case at realistic Reynolds numbers and to provide detailed information on the complex ow development and in particular the transition mechanisms. LES is less expensive and its application can be extended to higher Reynolds numbers. The test case considered is the T106 low-pressure cascade studied experimentally in various laboratories (Stadtmller, 2001; Hilgenfeld et al., 2002; Stieger et al., 2003). The cascade geometry is given in Fig. 1 with further details provided in Michelassi et al. (2002, 2003). In the experiments, the blades were assembled in a linear test rig and the wakes were generated by circular cylinders moving in front of the linear cascade. In the DNS and LES only one cascade channel could be afforded to be calculated, with periodic conditions on the pitch-wise boundaries. Such conditions can only be used when the cylinderto-blade-pitch ratio is an integer number. Unfortunately, in most of the experiments this ratio was not an integer number so that only few comparisons with the experimental results could be made. Two cases were studied: one is at a relatively low Reynolds number of Re = Uinf Cax / = 51800 (i.e. based on the axial 1 chord and inlet velocity) and a ratio of cylinder-to-blade-pitch of 2 . For this some limited measurements of Stadtmller (2001) are available. The other case is for a higher Reynolds number of Re = 148 000 and a cylinder-to-blade-pitch ratio of 1. For this situation, Wu and Durbin (2001) have carried out a

db

y 1.5 h

1 l 0.5

lb

0 1 -0.5 0 C Cax 1 2 2 x

Fig. 1. Geometry of T106 low-pressure cascade and grid used for LES of high-Reynolds number case (1 out of 12 nodes shown in both streamwise and pitchwise directions, from Michelassi et al., 2003).

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

149

1.2 1.0 0.8 0.6 Cp 0.4 0.2 0.0 -0.2 -0.4 -0.6 0.0 0.2 0.4 0.6 X/Cax 0.8 1.0
experiments DNS (in=45.5) LES (in=45.5) RANS (in=43.0)

1 0 -1 Cp -2 -3 -4 0.0
LES DNS refined [26] DNS coarse [26]

0.2

0.4 X/Cax

0.6

0.8

1.0

Fig. 2. Time-averaged static pressure coefcient distribution around T106 blade; left: low-Re case (from Michelassi et al., 2002); right: high-Re case (from Michelassi et al., 2003, DNS from Wu and Durbin, 2001).

DNS and this is taken for comparison. In both cases, the calculation domain is as shown in Fig. 1 (but only one of the two cascade channels shown there). For the low-Re case both DNS and LES have been carried out. In the DNS, the span-wise extent of the calculation domain was h = 0.2Cax and the grid used was 694 224 64, i.e. approximately 10 million points with the dimensionless wall distance of + the rst grid point y1 < 1. In the LES calculations, the span-wise extent was h = 0.15Cax and the grid + had 448 144 32 points, i.e. approximately altogether 2 million points and y1 was smaller than 2. In the high-Re case only an LES was carried out which was compared with the DNS of Wu and Durbin (2001) who used approximately 57 million grid points. In the LES reported here, h was 0.15Cax and a grid of 664 256 54 was used, i.e. altogether 10 million grid points. This grid is shown in Fig. + 1. The near-wall mesh sizes were x + < 70, z+ < 1520 and y1 = 13. All simulations enforce a no slip boundary condition on the blade and periodic conditions on the periodic boundaries in pitch-wise direction as well as in the span-wise direction. The inow boundary condition for the wakes is enforced by using the data base which was kindly made available by Wu and Durbin (2001), who generated the incoming wakes with preliminary LES. For the low-Re calculations, the details and a full discussion of the results are given in Michelassi et al. (2002) and Wissink (2003) and for the high Reynolds number LES calculations in Michelassi et al. (2003). We rst look at the pressure distribution along the suction and pressure surface of the blade which is given in form of the time-averaged static pressure coefcient Cp in Fig. 2. For the low-Re case, DNS and LES are almost indistinguishable and on the pressure side there is good agreement with the experiments of Stadtmller (2001) showing nearly constant pressure over the initial 60% of the blade and then fairly strong acceleration towards the trailing edge. On the suction side there is some deviation from the experiments which is due to compressibility effects in these which were not accounted for in the incompressible ow calculations and also due to some uncertainty about the inlet ow angle 1 . On the suction side there is rst a short region with deceleration and then acceleration until about x/Cax = 0.6 and then again an adverse pressure gradient region with deceleration towards the trailing edge leading to the intermittent occurrence of a separation bubble as will be shown shortly. In the high-Re case (right

150

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

Fig. 3. Phase averaged uctuating kinetic energy k at various phases from DNS of low-Re case, k increases with darkness (from Wissink, 2003).

part of Fig. 2), the pressure distribution along the two blade sides is similar and there is good agreement between the LES and the DNS of Wu and Durbin (2001). Fig. 3 shows snapshots of the phase-averaged uctuating kinetic energy calculated by DNS for the lower-Re case at different phases of the unsteadiness cycle determined by the movement of the wakegenerating cylinders. The gure illustrates how the wakes generated by the cylinders move through the cascade channel, their impingement on the blade and their interaction with the boundary layers. At the phase angle
= 5 6 it can be seen that the wake impinges twice on the suction side, once near the leading edge and almost simultaneously a previous wake near x/Cax = 0.8, the latter causing a suppression of the separation bubble which occurs near the trailing edge during times when no wakes impinge. Also noticeable is the production of uctuating kinetic energy in the middle of the cascade channel as the apex of the deformed wake becomes more and more pronounced. This production is due to straining of the wake as discussed in Wu and Durbin (2001). Fig. 4 gives two snapshots of 3D vortical structures in the wakes and near the suction side visualised by plotting iso-surfaces of the negative value of 2 , where 2 is the second largest eigenvalue of S 2 + 2 ; S is the symmetric part and the anti-symmetric part of the velocity gradient tensor (see Jeong and Hussain, 1995). The gure shows that accelerating ow along the upstream half of the suction side stretches the vortices inside the wake, resulting in elongated vortex structures. In the upper region, the wake is convected at a velocity smaller than the approach ow velocity and as a result the vorticity accumulates in the apex of the wakes. This accumulation, which can be seen in Fig. 4, coincides with the increase in

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

151

Fig. 4. Three-dimensional vortical structures near suction side of T106 blade (iso 2 contours) from DNS of low-Re case, at two times (T = wake passing period, from Wissink, 2003).

kinetic energy observed in Fig. 3. In the middle region, the disturbances on the suction side are suppressed by the ow acceleration, but near the trailing edge the presence of a multitude of tiny vortical structures indicates that there the ow is turbulent at all times and most of this turbulence is generated in a separation bubble which is intermittently present as will be discussed further below. Fig. 5 shows the vortex structures on the pressure side as calculated by LES for the high-Re casea similar behaviour was found also for the low-Re case and in the DNS of Wu and Durbin (2001). These authors argued that the elongated vortices appearing are not Grtler vortices due to an instability of the boundary layer on the concavely curved surface, but that these ow structures stem from the incoming vorticity of the wake. In the various background papers, these vortices are studied further by looking at uctuating velocity vectors in planes perpendicular to the pressure surface at various streamwise locations. There is generally good agreement about these structures between the LES and DNS. The vortices observed do not manage to trigger turbulence so that the boundary layer remains essentially laminar along the entire pressure surface as will be seen from the shape factor plotted in Fig. 8. The structures near the suction surface shown for the low-Re case in Fig. 6 at various times resulting from the LES are very different. In a plane perpendicular to the surface the span-wise vorticity is given in order to indicate the position of the wake in the vane. Near the leading edge, where the wakes impinge, disturbances are present, but these disappear in the acceleration region and are also not present near the leading edge at times without wake impingement. There is no signicant turbulence activity upstream of x/Cax 0.8. Near the trailing edge, again the small-scale turbulence seen already in the DNS results in Fig. 4 can be observed which stems from transition in an intermittent separation bubble as mentioned already. The difference in the boundary layer development on the suction and pressure sides is also very clear from Fig. 7 which shows the stream traces on the surfaces calculated by LES for the low-Re case. On the suction side, the traces are orderly and aligned with the ow direction and there are no signicant disturbances until x/Cax 0.85. From there on the traces are chaotic due to transition and turbulence induced by a local separation bubble. On the pressure side, the large and elongated ow structures that can be seen in Fig. 5 and have been observed and discussed already by Wu and Durbin (2001) give rise to the

152

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

2 4 6 Fig. 5. Instantaneous 2 surfaces near pressure side of T106 blade at four selected times t/T = 0 8 , 8 , 8 , 8 (from top to bottom, T = wake passing period) from LES of high-Re case (from Michelassi et al., 2003).

peculiar plot shown in the lower part of Fig. 7. This gure shows evidence of some interaction between the counter-rotating tube-like structures, which provoke local accumulation or defect of momentum in the span-wise direction due to a moderate span-wise migration of the axial vortices. The behaviour is not chaotic, conrming the observation that the boundary layer on the pressure surface does not become turbulent. This nding is supported by Fig. 8, in which the distributions of the average shape factor H and friction coefcient Cf are plotted along the pressure surface. If we exclude the immediate vicinity of the stagnation point, the shape factor H always remains above a value of around two typical for laminar

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

153

Fig. 6. Instantaneous iso-surface of axial vorticity on suction side of T106 blade from LES of low-Re case at times t/T = 0.2, 0.4, 0.7, 0.9 (from top leftclockwise, T = wake passing period). The blade-normal plane shows the spanwise vorticity.

0.1

0.25

0.5 x

0.75 Separation point

0.1

0.1 Stagnation point

0.2

0.3

0.4

0.5 x

0.6

0.7

0.8

0.9

Fig. 7. Instantaneous stream traces on suction surface (top, y + 25) and pressure surface (bottom, y + 13) of T106 blade from LES of low-Re case; vertical direction is blade span (from Michelassi et al., 2002).

boundary layers. There is a tendency for the average value to decrease in the middle part due to the effect of the impinging wakes until x/Cax 0.6, and at this location the phase averaged shape factor goes temporarily down to 1.6 as shown in the lower part of the gure. This indicates that at times when the wake passes the boundary layer is at the verge of becoming turbulent, but it recovers in between and on

154
5 4 3

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173


5

Cf H

4 3 2 1 0 1.0

2 1 0 0.0

0.2

0.4

0.6

0.8

x/Cax
2.8 2.6 2.4 2.2 2.0 1.8 1.6 1.4 0.0
2.8 2.6 2.4 2.2 2.0 1.8 1.6 20% Cax 1.4 1.0 1.5 2.0 0.0

60% Cax
0.5 1.0 1.5 2.0

0.5

2.8 2.6 2.4 2.2 2.0 1.8 1.6 1.4 0.0

<H>

0.5

1.0

Cf x 200
80% Cax
1.5 2.0

Fig. 8. Shape factor H and friction coefcient Cf on pressure side of T106 blade from LES of high-Re case; upper part: time-averaged values; lower part: phase-averaged values at three axial stations (from Michelassi et al., 2003).

average does not become turbulent. The lower part of Fig. 8 shows that the shape factor has an intermittent behaviour due to the impinging wakes only in the initial region x/Cax < 0.6 where the pressure is nearly constant, while in the subsequent region with strong acceleration the wakes have no effect and the shape factor is nearly constant at all times. The upper part of the gure shows that for x/Cax > 0.1 the average skin friction coefcient grows due to the ow acceleration which is mild in the rst 60% of the blade and quite strong further down stream. The values of the friction coefcient are always in the range of laminar or early transitional boundary layer on a at plate. Turning now to the suction side, Fig. 9 shows DNS results obtained for the low-Re case. In the left part, in an x
diagram the streamwise pressure gradient is given as well as lines of zero friction velocity which identify the separation regions. At the leading-edge where an adverse pressure gradient prevails, a small separation bubble centred at x/Cax 0.03 can be detected. This bubble remains present irrespective of the phase. The favourable pressure gradient downstream of x/Cax = 0.07 damps all disturbances originating from the unstable leading-edge separation bubble. Then beyond x/Cax = 0.6, the pressure gradient is again adverse and causes ow separation at (
, x/Cax ) (0.26, 0.85). At later times in the unsteadiness cycle, the bubble breaks up into two rolls that move downstream and disappear before
= 0.96. In the right part of the gure showing the vorticity near the trailing edge at
= 0.883, the development at these two rolls can be seen. Then for later times, the disturbances due to the impinging wakes are so strong that the separation is suppressed and does not exist in the rst quarter of the unsteadiness cycle. Turbulence

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

155

Fig. 9. Contours of dimensionless streamwise pressure gradient (Dp) and (white) lines, where U = 0 (left) and vorticity (vort) contours near trailing edge (right) on suction side of T106 blade for low-Re case from DNS (from Wissink, 2003).

0.020 DNS 0.015 LES

Cf

0.010

0.005

0.000 0.0 0.2 0.4 x/Cax 0.6 0.8 1.0

Fig. 10. Time-average friction coefcient along suction side of T106 blade for high-Re case (from Michelassi et al., 2003, with DNS from Wu and Durbin, 2001).

is produced in the separation-bubble region (see Wissink, 2003), and the ow is transitional to turbulent at all times near the trailing edge. In contrast to the low-Re case, the boundary layer does not separate on the suction side in the high-Re case. This can be seen from the average friction coefcient Cf calculated by LES and DNS along this surface shown in Fig. 10. Cf decreases strongly in the adverse pressuregradient region where the ow is still laminar, but it does not reach zero as transition occurs and causes it to increase again strongly. This

156

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

t/T = 15.00

t/T = 15.25

x/Cax=0.6

t/T = 15.50

t/T = 15.75

Fig. 11. Iso lines of vertical velocity component over suction side surface of T106 blade from LES for high-Re case. The grey scale on the back plane corresponds to the uctuating velocity and is added to identify the location of the wakes (from Michelassi et al., 2003).

behaviour is predicted in good qualitative agreement between DNS and LES, and also the quantitative agreement is good until transition occurs. LES predicts transition to happen a little later than DNS and hence also Cf rises further downstream. It was concluded in Michelassi et al. (2003) that LES does not reproduce transition as accurately as DNS due to the lower resolution of the smaller-scale processes. Finally, Fig. 11 shows uctuations predicted by LES on the suction surface for the higher-Re case. The back plane perpendicular to the surface serves to identify the instantaneous position of the wake. Near the leading edge, the effect of the wakes is much smaller than in the DNS of Wu and Durbin (2001). Any small disturbances introduced do not survive and are damped out entirely by the favourable pressuregradient downstream of x/Cax 0.6. The DNS indicates the presence of turbulent spots which grow and then cause transition at x/Cax 0.76, while in the LES the perturbations observed in this region are too smooth to correspond to turbulent spots. The difference is likely to be due to the poorer resolution in the LES but may also have to do with the wake decaying more on its travel through the cascade due to a relatively coarse grid used in LES in the middle of the cascade channel. Around x/Cax 0.8, the disturbances also grow in LES rapidly and transition occurs so that the trailing edge region is then fully turbulent. The DNS and LES of the ow in low-pressure turbine cascades have shown that the various complex phenomena occurring in these such as the development of vortex structures, transition and also intermittent

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

157

separation in the case of low Reynolds number can be captured by both methods. Realistic calculations are therefore possible with these methods for Reynolds numbers of practical relevance (here up to 150 000, based on inlet velocity and axial chord) but the computing times required are fairly large, approximately 10 times higher for DNS than for LES. The cheaper LES deliver less detail of the turbulent ow structures, but there was a good agreement between the two calculations for the general ow patterns. For the low-Re case even the quantitative agreement was fairly good while in the higher-Re case the complex mechanisms that trigger transition in attached boundary layers could not be fully reproduced by LES which must hence be considered not a very suitable tool for studying such transition mechanisms.

4. Flow around the Ahmed car body The next example concerns the ow around a generic car body which in spite of its relatively simple geometry has many features of the complex, fully 3D ow around real cars. The so-called Ahmed body was chosen as test case. This body was rst dened by Ahmed et al. (1984) and its geometry is given in Fig. 12. The body is near-rectangular with a rounded front part and has a slant back. Ahmed et al. (1984) have carried out experiments with several slant angles and described the characteristics of the ow for various angles. They found that most of the drag of the body is due to pressure drag, which is generated at the rear end. This can be seen from Fig. 13, which on the right gives a reproduction of their plot of the various contributions to the drag coefcient as a function of the slant angle. (CR is the contribution from friction). The maximum drag was found for a critical slant angle of 30 . Above this angle, a sudden drop in drag occurs which corresponds to a drastic change in the ow in the wake. Below this angle, strong counter-rotating vortices emanating from the sloping edges of the body are present and the ow separates in the middle region of the top edge and reattaches at the sloping surface. For angles above the critical angle, the counter-rotating vortices are much weaker, the separation occurs along the entire top and the side edges, and there is no reattachment on the sloping surface. The differences

Fig. 12. Geometry of Ahmed body model, dimensions are in millimeter (from Hinterberger et al., 2004).

158

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

Fig. 13. Development of the ow in the rear of the Ahmed body for two slant angles (left, courtesy of S. Becker and H. Lienhart, LSTM Erlangen) and dependance of drag coefcient on slant angle (right; from Ahmed et al., 1984).

are illustrated by the sketches given in the left of Fig. 13 and also by the ow visualisation pictures in Fig. 14 from a later experiment of Lienhart et al. (2000) which show oil-ow pictures for the sloping surface and visualisation of the wake ow with smoke injection for the slant angles 25 and 35 , i.e. just below and above the critical angle. These more recent measurements of Lienhart et al. (2000) were performed using the same body, but at a somewhat lower velocity than in the experiments of Ahmed et al. (1984). Lienhart et al. (2000) also measured the mean and uctuating velocities by LDA which will be taken for comparison with the LES presented in this paper. It should be noted that this ow was also a test case at two workshops (Jakirlic et al., 2001; Manceau and Bonnet, 2002)and the results presented there have shown that this is indeed a challenging ow for which it is difcult to obtain accurate predictions of mean-velocity and turbulence-intensity proles with classical RANS methods. LES is presented here for the case with = 25 as studied experimentally by Lienhart et al. (2000), and the full details are described in Hinterberger et al. (2004). Work is in progress in the authors group on calculations for the case with = 35 . The ow geometry is as shown in Fig. 12 and the Reynolds number is Re = 2.8 106 (based on the length of the body). The inow plane of the calculation domain was placed at a distance of 1.3 body lengths upstream of the body. This corresponds roughly to ve body heights and is a distance where blocking effects can be expected to disappear. At the inow section, a uniform axial velocity prole is imposed. The width of the calculation domain was ve body widths and

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

159

Fig. 14. Oil ow visualisation on slant surface of Ahmed body and visualisation of ow in wake for two slant angles (left: 25 , right 35 , courtesy of S. Becker and H. Lienhart, LSTM Erlangen).

the height ve body heights, analogous to the experiment. At both side boundaries and at the top boundary, free-slip conditions are applied. The outow plane of the domain was placed ve body lengths behind the body to ensure that the outow condition does not affect the near-body wake. A convective boundary condition was used for the velocities there. Finally, wall functions are used because of the high Reynolds number which does not allow a wall-resolving LES and even with this approach the resolution was not sufcient. The wall function used is similar to the Werner and Wengle (1993) approach, but assuming an instantaneous logarithmic prole instead of a power-law prole. It is applied at the walls of the vehicle and at the bottom of the channel. A Smagorinsky SGS model was employed in the LES with a model constant of Cs = 0.13. The calculations were performed on two multi-block grids. The coarser one consists of 93 blocks and 8.8 106 cells and the ner one of 214 blocks and 18.5 106 cells. The ne grid is shown in Fig. 15 indicating clearly the renement near the body. For both grids, the near-wall cell centre has a wall distance on average of about 40 wall units, but varies from approximately 10 in the separated regions along the slant back to 150 close to the top rear edge. The span-wise and stream-wise extent of the grid cells is up to a factor of 10 larger for the ne grid. This factor is even larger for the coarse grid. This means that even in connection with using wall functions, the boundary layer developing on the body surface is highly under-resolved in spite of the many grid points employed. Fig. 16 presents the calculated pressure distribution around the Ahmed body. At the front face the stagnation pressure prevails and there is a clear excess pressure in the entire front part. In the rear, the under pressure due to the ow separation in this region can be seen. Integrating the pressure and friction forces in the x direction yields a drag coefcient of cD = 0.3 which compares to the value of 0.29 measured by Ahmed et al. (1984), albeit at a somewhat higher Reynolds number. As found by Ahmed et al., the contribution of the pressure drag is about 80% and the friction drag only about 20%.

160

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

Fig. 15. Fine grid (18.5 106 ) cells for calculation of ow around Ahmed body.

The mean ow around the front of the body is very well predicted, as was shown in Hinterberger et al. (2004). In Figs. 17 and 18, the mean streamwise velocity proles and the root mean square velocity uctuations in the symmetry plane are compared with the experimental data in the rear part of the body and the near wake. Calculations obtained with both coarse and ne grid are included. The general agreement with the experiment is quite reasonable taking into account that neither grid is ne enough to resolve adequately the boundary layer development on the body up to the slant back. However, there are some signicant discrepancies between the computations and the experiments mainly concerning the velocity proles on the slant back. In the experiment, the ow separates right at the corner of the sloping surface and reattaches in the middle of that surface. In the simulations, the ow rst stays attached before it separates somewhat downstream of the corner and it does not reattach on the slant back. These discrepancies are most likely due to the poor resolution of both grids near the wall leading to an incorrect prediction of the boundary layer approaching the slant back. The main differences between the results of the two simulations are found near the beginning of the sloping surface. There, the ne grid yields turbulence intensities which are very close to the experimental ones while for the coarse grid they are too low (Fig. 18). In addition, the prediction of the separation point is closer to the experiment in the simulations performed with the ner grid, in which the ow separates earlier (Fig. 17). Fig. 19 compares the calculated and measured mean secondary-ow velocity vectors at two transverse planes, one being at the rear end of the body (x = 0) and the second 200 mm down the rear end. It can be seen that the development of the counter-rotating vortices is predicted in close agreement with the

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

161

Fig. 16. Calculated pressure distribution for ow around Ahmed body.

experiments. These vortices develop already halfway downstream of the slant back and grow while they approach the end of the body and decay only further downstream in the wake. Hinterberger et al. (2004) have shown in addition that also the turbulence level in these vortices is predicted correctly. The calculation results show complex time-dependent ow features in the wake region. In Fig. 20, two typical instantaneous velocity elds in the symmetry plane are illustrated. In the left part, an instantaneous eld is captured in which the ow tends to reattach on the slant back as was observed on average in the experiments. However, in the right part of the gure the ow in that zone is completely separated, and this state occurs more often than reattachment so that on average the calculations show no reattachment. The unsteadiness of the near-body wake is well captured in the simulations. The ow between the body and the ground plate has a rather strong inuence on the shape of the recirculation zone, and also here an intermittent behaviour can be seen in which the recirculation zone is sometimes shorter (left picture) and sometimes longer (right picture). Fig. 21 illustrates the presence of large-scale vortices in the recirculation zone by showing the isosurfaces of pressure uctuations at one instant (a sequence is given in Hinterberger et al., 2004). The main part gives a perspective view, while the three pictures on the left give the side view, top view and view from the back, respectively. It can be seen that large span-wise vortex tubes are generated near the leading edge of the slanted surface, which grow and roll down that surface. Furthermore, vortex tubes that wrap the counter-rotating vortices coming off the slant side edges can be detected. Animations have shown that these vortex tubes form helical structures as they move downstream.

162

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

Fig. 17. Mean streamwise velocity proles in rear of Ahmed body and near-wake (in symmetry plane, from Hinterberger et al., 2004).

In summary, the ow around the Ahmed body with slant back is a very challenging problem due to the complex 3D unsteady ow behaviour in the rear of the body and the wake and due to the high Reynolds number. The LES has managed to simulate the overall ow behaviour quite well, capturing the complex structures developing in the rear. However, because of the high Reynolds number the LES could not resolve properly the development of the boundary layer along the body which was considered the main reason for the observed discrepancies with experiments, in particular the failure to predict reattachment in the mean of the ow on the sloping surface. It appears that for such a high Reynolds number situation with large regions of thin attached boundary layers on the body a detached eddy simulation (DES) approach or in general a coupled LES/RANS calculation would be more suitable in which the development of the attached boundary layer is simulated by the RANS mode. In any case, further calculations, also for different slant angles, should be carried out for this challenging test problem. 5. Flow past a nite-height cylinder The last example concerns the ow past a circular cylinder of nite height mounted on a at plate. Such ow situations occur in practice in the ow around towers, stacks or round buildings, and nite-height cylinders are also of relevance as simplied roughness or vegetation elements in atmospheric boundary layers, open-channel or river ows. A complex 3D ow develops in such a situation as can be seen from the oil-ow picture of Hlscher (1993) reproduced in Fig. 22. The ow is particularly complex around

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

163

Fig. 18. Root mean square streamwise velocity uctuations in rear of Ahmed body and near-wake (in symmetry plane, from Hinterberger et al., 2004).

the top but also near the bottom where the ow around the cylinder interacts with the boundary layer on the ground plate forming a horseshoe vortex. The ow past long circular cylinders is characterised by separation in the rear with vortex shedding and the formation of a von Krmn vortex street and was much studied. End effects on the top and the ground plates alter this behaviour substantially. The ow behaviour is sketched for cylinders of different height-to-diameter ratio H /D in Fig. 23. When this ratio is reduced successively, the regular alternating vortex shedding, which is typical for a long cylinder, is more and more replaced by symmetrically shed vortices in the range H /D 62 (Kawamura et al., 1984; Kappler, 2002). Vortex shedding is mostly suppressed for values around and below H /D 2 (Okamoto and Yagita, 1973; Kappler, 2002). In addition to the height-to-diameter ratio H /D and of course the Reynolds number, the thickness of the approaching boundary layer relative to the cylinder height H has also a strong inuence on the ow development, as addressed by Kawamura et al. (1984). The conguration simulated by LES is sketched in Fig. 24. It was studied experimentally in a water tunnel using ow visualisation and laser doppler anemometry by Kappler (2002). The height-to-diameter ratio is H /D = 2.5 and the relative approach ow boundary layer thickness /H = 0.1, i.e. the cylinder was mostly exposed to uniform approach ow. The Reynolds number Re = U D/ is 43 000, so that the ow is in the subcritical regime in which a laminar boundary layer develops along the cylinder wall and transition to turbulence occurs shortly after separation through a KelvinHelmholtz instability and further span-wise instabilities.

164

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

Fig. 19. Secondary mean-ow velocity vectors in two transverse (y z) planes at rear end and in wake of Ahmed body (from Hinterberger et al., 2004).

Fig. 20. Instantaneous streamwise velocity elds in the symmetry plane of Ahmed body at 2 times T, Ub is the bulk velocity (from Hinterberger et al., 2004).

In the experiment, the width of the tunnel was 7D and the height 5D introducing a blockage of 7.3%. The width and the height of the computational domain were selected to be the same, hence introducing the same blockage as in the experiment. The inow boundary was located 7.5D upstream of the centre of the cylinder, imposing a constant inow velocity U . The outow boundary was located 12.5D downstream

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

165

Fig. 21. Contours of iso-surface of pressure uctuations in rear part of Ahmed body.

Fig. 22. Oil-ow picture for ow around surface mounted nite-height circular cylinder (from Hlscher, 1993).

166

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

Fig. 23. Sketches of the ow eld around a cylinder of nite-height; left: situation where cylinder is longer than the critical length for vortex shedding; right: same for a shorter cylinder (from Kawamura et al., 1984).

Fig. 24. Finite-height cylinder conguration studied by LES.

of the cylinder centre and a convective boundary condition was imposed there. The boundary condition on the cylinder surface was a no-slip condition and the conditions at the top and at the side walls were free-slip. On the solid ground plate, the wall function due to Werner and Wengle (1993) was used. The calculation domain was discretised with 6.4 million grid points and a two-dimensional cut in wall parallel direction through the grid is shown in Fig. 25. A full description of the work is given in Frhlich and Rodi (2004). Fig. 26 shows a snap shot of iso-surfaces of instantaneous pressure deviation (from the average), giving an impression of the instantaneous ow and the structures developing. They show that the separation at the sharp front corner of the cylinder top is fairly regular with lateral vortex rollers which interwind and

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

167

Fig. 25. Two-dimensional cut in wall parallel direction through grid used for LES of ow around nite-height cylinder (from Frhlich and Rodi, 2004).

Fig. 26. Iso-surface of instantaneous pressure deviation for ow around a nite-height cylinder; viewed from the top (right) and from an oblique angle from the rear (left) (from Frhlich and Rodi, 2004).

Fig. 27. Instantaneous picture of vortex shedding close to the ground plate. Left: visualisation by means of a tracer in the experiment (Kappler, 2002); right: instantaneous u velocity from LES (from Frhlich and Rodi, 2004).

168

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

Fig. 28. Temporal evolution of instantaneous lift (lower curve) and drag (upper curve) force on nite-height cylinder as calculated by LES (from Frhlich and Rodi, 2004).

Fig. 29. Vortex behind a nite-height cylinder; visualisation by means of stream ribbon and stream traces (left); average stream lines and contours of streamwise velocity U in plane 0.5D downstream of cylinder (right) (partly from Frhlich and Rodi, 2004).

merge upon travelling downstream along the roof. At the rear, in the top region, the separation process is highly complex and very irregular. Further down towards the bottom wall the structures become more regular as separation takes place in a more coherent way and larger vortices with axes parallel to the cylinder axes are formed. In the experiment, no regular vortex shedding was detected near the top, but regular alternating shedding near the bottom surface. Fig. 27 shows a tracer photograph in this region from the experiment and compares it with a visualisation picture from the calculations, showing that these indeed also yield alternating vortex shedding. Further downstream, in the wake, the shed vortices increase in size and become smoother. Highly distorted Krmn vortices are visible in Fig. 26. They are bent inwards and backwards due to the downward motion behind the cylinder. It is clear that there is no regular Krmn vortex street developing at this height-to-diameter ratio; regular Krmn vortices parallel to the cylinder axes appear only for longer cylinders.

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

169

Fig. 30. Surface streamlines on cylinder (upper part; left is oblique view with ow from left, right is rear view) and ground plate (lower part) as calculated by LES (from Frhlich and Rodi, 2004).

The irregular separation and shedding processes produce relatively irregular forces on the cylinder. This is illustrated in Fig. 28, showing the temporal evolution of the instantaneous lift and drag forces. While a dominant frequency f with a Strouhal number of about St = f D/U = 0.16 can be discerned, the lift coefcient exhibits an irregularly changing amplitude and can have an average different from zero over a certain number of shedding cycles. The trailing vortices sketched in Fig. 23 can also be detected from the LES results. Fig. 29 in the right part shows secondary ow stream traces at a distance of 0.5D past the rear end of the cylinder. Vortices can clearly be seen which are the tip vortices that have been swept downwards by the down ow behind the cylinder as seen from the streamlines shown later in Fig. 32. One of the tip vortices is also visualised in the left part of Fig. 29 by means of a stream ribbon. At a height of about y/D 1 this ribbon leaves the tip vortex and stretches downstream until the exit of the domain, still spiralling but at a much lower pace. Two further stream traces have also been introduced in close vicinity for clarication. The black one spirals in the interior of the ribbons trace and resembles the tip vortex in Fig. 23. The second, white one remains further outward close to the ribbon but continues down to the bottom plate.

170

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

Fig. 31. Average streamlines (left) and streamwise velocity uctuations (right) at height y/D = 1.5 for ow around a nite-height cylinder; top: experiment (Kappler, 2002), bottom: LES (from Frhlich and Rodi, 2004).

Fig. 30 shows surface streamlines along the cylinder surface and the bottom wall. The stagnation line is clearly visible in the left oblique view of the gure. Near the bottom, a saddle point exists at y/D = 0.2 resulting from the oncoming boundary layer. Near the top the stagnation line splits up into a fan of streamlines due to the end effect. The separation line along the side wall is also clearly visible. It is located at a fairly constant angle of 80 from the stagnation line but curves towards the rear near the bottom and the top. The ow in the rear of the separation line shows a substantial upward motion. Near the top, a focus is visible in the right picture resulting from the upward rear motion and the counter-acting downward motion in the centre plane. On the roof, the reattachment of the ow can be seen in the middle of the roof. The foci on the rear of the cylinder close to the top in the right part of the gure are the foot prints of the two tip vortices which separate from the cylinders free end and reach down into the wake as discussed before leading to the vortices seen in Fig. 29. On the ground plate, the mean recirculation behind the cylinder clearly leaves its foot print as seen in the lower part of Fig. 30. This gure also indicates that a small horseshoe vortex forms in front of the cylinder and bends around it. The recirculation region behind the cylinder can also be seen from the mean streamlines shown at a height of y/D = 1.5 in Fig. 31, giving a comparison of computed and measured results. The gure gives evidence that the size and form of the recirculation region is predicted very well by the LES. The gure also compares calculated and measured contours of the streamwise uctuations u . The agreement is again fairly good, with the maximum of the uctuations occurring in the shear layer bordering the separation region. The size of the zone with signicant uctuations is quite well predicted. Finally, in Fig. 32, mean streamlines in the vertical centre plane are compared between LES and experiments, showing the ow over the top of the cylinder and behind the cylinder. Again, the agreement is fairly good, showing clearly the recirculating motion in this plane behind the cylinder with a strong downward motion impinging on the ground plate, a reverse ow near the ground plate and an upward motion in the rear of the cylinder. The separation on the top can also be seen in the LES results, but this could not be resolved in the experiment. In the right part of the gure, contours of the vertical uctuations v shown for the LES calculations and

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

171

Fig. 32. Average streamlines (left) and vertical uctuations (right) in centre plane (z/D = 0) of ow past a nite-height cylinder; top: experiment (Kappler, 2002), bottom: LES (from Frhlich and Rodi, 2004).

the measurements, and again there is reasonably good agreement, with a maximum of the uctuations in the shear layer region bordering the recirculation zone. The LES of ow around a surface mounted nite-height circular cylinder have shown that, in comparison with experiments, the main features of this complex ow can be well predicted. Among such features are the details of the interaction of the separation from the sidewalls and the roof with the downwash behind the cylinder, the tip vortices, the vortex shedding near the ground and the horseshoe vortex developing there. In particular, the LES was capable of reproducing the various unsteady ow structures which are characteristic for this type of ow.

6. Conclusions In this paper, it was shown that DNS is possible for certain engineering ows at realistic Reynolds numbers, such as the ow through low-pressure turbine cascades, and that this method allows us to study all details of the complex ow phenomena prevailing such as transition and the travel of wakes through the cascade channels inuencing strongly the transition. However, such calculations are very expensive, running several months on the biggest supercomputers available. Hence, DNS is clearly not the method for routine engineering calculations, but it is an important tool for studying the details of certain complex

172

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

ow phenomena such as near-wall turbulence and transition mechanisms, and it will be used increasingly for this purpose. LES could also be applied successfully to such ows but also to geometrically more complex, fully 3D ow situations where this method was found to produce realistically the main features, including the unsteady behaviour but also the time-average quantities. LES is clearly superior to RANS whenever large-scale structures dominate the behaviour and when unsteady effects like vortex shedding are present. LES calculations are cheaper than DNS, but they are still fairly costly. However, they are often affordable on modern computers, for some problems even on clusters of PCs, and they will be used more and more also for practical applications in cases where the aforementioned effects are important. For high Reynolds number ows involving walls, some special near-wall modelling is necessary and here LES/RANS coupling of some kind will be the method of the future. The further increase in computer power will certainly lead also to an increased use and exploitation of DNS and LES. Acknowledgements The calculations reported here were sponsored by the German Research Foundation (DFG) through various research projects. The author is grateful to J. Frhlich, Ch. Hinterberger and J. Wissink for providing various gures on their calculation results included in this paper. References
Ahmed, S.R., RAM, G., Faltin, G., 1984. Some salient features of the time-averaged ground vehicle wake. SAE Paper No. 840300. Breuer, M., Rodi, W., 1996. Large eddy simulation of complex turbulent ows of practical interest. In: E. Hirschel (Ed.), Flow Simulation with High Performance Computers II, Notes on Numerical Fluid Mechanics, vol. 52, Vieweg, Braunschweig, pp. 258274. Frhlich, J., Rodi, W., 2004. LES of the ow around a circular cylinder of nite-height. Int. J. Heat Fluid Flow 25, 537548 (Figs. 2532 or parts thereof were reprinted from this publication with permission from Elsevier). Germano, M., Piomelli, U., Moin, P., Cabot, W., 1991. A dynamic subgrid-scale eddy-viscosity model. Phys. Fluids A 3, 17601765. Hilgenfeld, L., Stadtmller, P., Fottner, L., 2002. Experimental investigation of turbulence inuence of wake passing on the boundary layer development of highly loaded turbine cascade blades. Flow Turbul. Combust. 69, 229247. Hinterberger, C., Garcia-Villalba, M., Rodi, W., 2004. Large eddy simulation of the ow around the Ahmed body. In: McCallen R., et al. (Eds.), The Aerodynamics of Heavy Vehicles: Trucks, Buses and Trains, Lecture Notes in Applied and Computational Mechanics, vol. 19. Springer, Berlin. Hlscher, N., 1993. Ein multivarianter Ansatz fr die aerodynamische bertragungsfunktion der Winddrcke in atmosphrischer Grenzschichtstrmung. Ph.D. Thesis, Ruhr-Universitt Bochum, Germany. Jakirlic, S., Jester-Zrker, R., Tropea, C. (Eds.), 2001. Proceedings of the Ninth ERCOFTAC IAHR COST Workshop on Rened Turbulence Modelling, Technical University Darmstadt. Jeong, J., Hussain, F., 1995. On the identication of a vortex. J. Fluid Mech. 285, 6994. Kappler, M., 2002. Experimentelle Untersuchung der Umstrmung von Kreiszylindern mit ausgeprgt dreidimensionalen Effekten. Ph.D. Thesis, Institute for Hydromechanics, University of Karlsruhe. Kawamura, T., Hiwada, M., Hibono, T., Mabuchi, I., Kamuda, M., 1984. Flow around a nite circular cylinder on a at plate. Bull. JSME 27 (232), 21422151. Lienhart, H., Stoots, C., Becker, S., 2000. Flow and turbulence structures in the wake of a simplied car model (Ahmed model). In: DGLR Fachsymp. der AG STAB. Manceau, R., Bonnet, J.P. (Eds.), 2002. Proceedings of the Tenth ERCOFTAC IAHR QNETCFD Workshop on Rened Turbulence Modelling, University of Poitiers.

W. Rodi / Fluid Dynamics Research 38 (2006) 145 173

173

Michelassi, V., Wissink, J., Rodi, W., 2002. Analysis of DNS and LES of ow in a low pressure turbine cascade with incoming wakes and comparison with experiment. J. Flow Turbul. Combust. 69, 230295. Michelassi, V., Wissink, J.G., Frhlich, J., Rodi, W., 2003. Large eddy simulation of ow around a low pressure turbine blade with incoming wakes. AIAA Journal 41, 21432156 (Figs. 1, 2, 5, 8, 10, 11 or parts thereof were reprinted from this publication by permission of the American Institute of Aeronautics and Astronautics, Inc.). Okamoto, T., Yagita, M., 1973. The experimental investigation on the ow past a circular cylinder of nite length placed normal to the plane surface in a uniform stream. Bull. JSME 16 (95), 805814. Spalart, P.R., Jou, W.-H., Strelets, M., Allmaras, S.R., 1997. Comments on the feasibility of LES for wings and on a hybrid RANS/LES approach. In: Liu, C., Liu, Z. (Eds.), Advances in DNS/LES. Greyden Press, Columbus, OH. Stadtmller, P., 2001. Investigation of wake-induced transition on the LT turbine cascade T106 A-EIZ. DFG Verbundprojekt Fo 136/11, Version 1.0, University of the Armed Forces Munich, Germany. Stieger, R., Hollis, D., Hodson, H., 2003. Unsteady surface pressures due to wake induced transition in a laminar separation bubble on a LT turbine cascade. Paper ASME-GT 2003-38303. Werner, H., Wengle, H., 1993. Large eddy simulation of turbulent ow over and around a cube in a plane channel. In: Durst, F., et al. (Eds.), Turbulent Shear Flows, vol. 8. Springer, Berlin, pp. 155168. Wissink, J.G., 2003. DNS of separating low Reynolds number ow in a turbine cascade with incoming wakes. Int. J. Heat Fluid Flow 24, 626635 (Figs. 3, 4, 9 were reprinted from this publication with permission from Elsevier). Wu, X., Durbin, P.A., 2001. Evidence of longitudinal vortices evolved from distorted wakes in a turbine passage. J. Fluid Mech. 446, 199228.

Das könnte Ihnen auch gefallen