You are on page 1of 11

International Journal of Heat and Fluid Flow 43 (2013) 233–243

Contents lists available at SciVerse ScienceDirect

International Journal of Heat and Fluid Flow


journal homepage: www.elsevier.com/locate/ijhff

Aerodynamic characterization of the wake of an isolated rolling wheel


Emma Croner a,1, Hervé Bézard b,⇑, Christophe Sicot c, Guillaume Mothay c
a
MICHELIN, Centre de Technologie Ladoux, Clermont-Ferrand Cedex 63040, France
b
ONERA, The French Aerospace Lab, F-31055 Toulouse, France
c
Institut Pprime, UPR CNRS 3346, ENSMA, Université de Poitiers, France

a r t i c l e i n f o a b s t r a c t

Article history: The flow around the wheels of road vehicles exhibits complex unsteady 3D phenomena, including mas-
Available online 31 May 2013 sive boundary layer separations, recirculation areas and 3D vortical structures. Although the mechanisms
generating the wake main vortices have been identified, their exact topology is still indeterminate and a
Keywords: description of their unsteady evolution has not yet been proposed.
Aerodynamics The aim of this paper is to characterize the structure and unsteady motion of an isolated wheel wake.
CFD For this purpose, wind tunnel investigations have been carried out with a smooth rotating wheel in con-
Hot-wire
tact with a moving belt and compared to URANS simulations based on the geometry used for experi-
PIV
URANS
ments.
Wheel CFD calculations appeared to have successfully predicted the main features of the flow. Particle Image
Velocimetry and hot-wire anemometry measurements coupled with numerical simulation results
enabled the origin and the nature of the main vortical structures in the near-wake to be confirmed
and their unsteady behaviour to be investigated. The results show good agreement with previous knowl-
edge and provide new insight into the unsteady features of the wake.
Ó 2013 Elsevier Inc. All rights reserved.

1. Introduction nomenon, confirmed by smoke visualizations, is the boundary


layer separation occurring further upstream on the top of a rotat-
Car wheels are responsible for 30–40% of the total drag on mod- ing wheel compared to a stationary one because of the fluid
ern vehicles (Elofsson and Bannister, 2002; Régert and Lajos, 2007). entrainment by the wall. The wall motion also causes the separa-
In addition to the own drag of these bluff bodies, this high contri- tion to be located above the surface, with a specific saddle point
bution is also due to the strong interactions between their wake topology. Added to the jetting-vortices effects, the upstream
and the vehicle underbody through pressure modifications. Hence, boundary layer separation causes the wake of the rotating wheel
as body shape optimization reaches its limits, wheels aerodynam- to be taller and narrower than the one of the stationary wheel.
ics is becoming a major concern in order to reduce fuel consump- This pressure distribution analysis was completed by Mears
tion and it is therefore essential to gain understanding of the wake et al. (2002) who reproduced Fackrell’s measurements with mod-
dynamics. ern devices, thus confirming previous observations and revealing
Fackrell and Harvey (1974) carried out the first experimental the negative pressure peak already suggested.
work about an isolated wheel in realistic conditions, namely a Most of the following studies emphasized the importance of
rotating wheel in contact with a moving floor. Pressure distribution reproducing realistic rolling conditions to analyse vehicle aerody-
measured on the wheel surface highlighted two main specificities namics. Cogotti (1983) and Mercker and Berneburg (1992) outlined
of the flow. First, upstream of the contact area, the rotating wheel the importance of the wheel and ground motions to ensure correct
and the moving ground behave like a viscous pump and cause wake development and consequently to estimate properly the
pressure coefficients to rise above 1. A lateral jet stream, referred aerodynamic coefficients. Contact between the moving surfaces
as jetting, is produced to each side of the wheel generating two is indeed the only way to generate the flow features mentioned
near-ground vortices. A negative pressure peak downstream of above. These considerations on isolated wheels were extended to
the contact area was predicted but not observed. The second phe- full vehicle configurations by experiments, summarized by Elofs-
son and Bannister (2002). All authors highlighted the major effect
⇑ Corresponding author. of moving ground and rotating wheels on the vehicle’s total lift and
E-mail addresses: emma.croner@onera.fr (E. Croner), herve.bezard@onera.fr drag coefficients. In addition, Mlinaric and Sebben (2008) empha-
(H. Bézard), christophe.sicot@lea.ensma.fr (C. Sicot), guillaume.mothay@etu. sized the importance of the contact patch dimensions, especially
ensma.fr (G. Mothay). the transversal one.
1
Tel.: +33 5 62 25 28 28; fax: +33 5 62 25 25 83.

0142-727X/$ - see front matter Ó 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.ijheatfluidflow.2013.04.008
234 E. Croner et al. / International Journal of Heat and Fluid Flow 43 (2013) 233–243

Nomenclature

C force coefficient (–) Subscripts and Superscripts


D wheel diameter (m) i at the inlet of the test section
f frequency (Hz) L, D, S lift, drag, side force
L wheel width (m) P in-plane
V mean velocity vector norm (m/s) x, y, z axial, transversal, vertical

From a more general point of view, the authors’ main objectives downstream from the top of the wheel instead of two contra-rota-
were to quantify the impact of different elements on the vehicle’s tive vortices. Wäschle (2007) described the flow topology with a
global aerodynamic performances in order to discover whether or ring vortex on the top and a wake-horseshoe vortex downstream
not it is necessary to take them into account during measurements from the contact area. He considers that this structure dominates
or numerical simulations. Hence, it was found that wheel charac- the wake: the massive vortices in the lower part of the wake result
teristics such as grooves or deformation had a major effect on from the merging of the weak jetting-vortices with the powerful
the car aerodynamics through underbody and base pressure mod- wake-horseshoe vortex. Finally, low-Reynolds Direct Numerical
ifications (Modlinger et al., 2007). All experimental work, con- Simulations (DNS) by Pirozzoli et al. (2012) highlighted a vortex
ducted since the 1970s, as well as the numerical work during the pair detaching from the rear part of the tyre between the upper
last decade, concluded that the wheels had a major impact on vortices and the near-ground vortices.
car aerodynamics and consequently outlined the importance of Although the main flow features have already been studied
understanding the flow that develops around them. both experimentally and numerically, authors do not yet agree
From consideration of previous works and vortex theory, Merc- on the exact flow topology surrounding a rotating wheel. More-
ker and Berneburg (1992) presented a schematic picture of the over, few numerical studies were conducted with unsteady calcu-
near-wake vortical structures including the two jetting-vortices, lations and, to the author’s knowledge, none of them tried to
two vortices downstream of the upper wheel boundary separation characterize the unsteady behaviour of the wake. Only Basara
– specific to lift bodies – and two weak central vortices created at et al. (2000) suggested the presence of a vortex shedding phenom-
the centres of the rims. More recently, Saddington et al. (2007) enon and some authors (Damiani et al., 2004; Krajnović et al.,
revisited this model in the light of Laser Doppler Anemometry 2011) evaluated the drag evolution of a wheel in a wheelhouse.
(LDV) measurements: the central vortices have not been observed However, for bluff bodies, the use of unsteady calculation is known
and are supposed to be quickly swept up into the lower and upper to provide more accurate solutions for both the spatial features of
ones; moreover the upper vortices appeared to be weaker and to the flow and the aerodynamic forces acting on the body (Iaccarino
merge with the jetting-vortices within one diameter downstream and Durbin, 2000). Unsteady features can also ease the under-
of the wheel axis (see Fig. 1). standing of the mechanisms generating the flow structures. This
The use of Computational Fluid Dynamics (CFD) added a new paper aims to complete current knowledge by analyzing the wake
insight into the understanding of car wheel aerodynamics. The first structure topology and unsteadiness in order to better understand
simulations with isolated wheels (Basara et al., 2000; Skea et al., the main vortex dynamics and their impact on aerodynamic forces.
2000) tended essentially to demonstrate the CFD ability to predict
specific flow features through comparisons of pressure coefficients. 2. Experimental settings
Besides parametric studies, the following works provided detailed
visualizations of the flow around isolated wheels and wheels The wind tunnel tests campaign took place in the Aero Concept
inserted in wheelhouses and suggested different sketches of the Engineering (ACE) facility in Magny-Cours, France. Particle Image
main flow structures. With regard to the rotating wheel, Velocimetry (PIV) and hot-wire anemometry measurements were
McManus and Zhang (2006) observed an arch-shaped vortex just

Fig. 1. Sketch of the flow vortical structures in the wake of an isolated wheel by
Saddington (2007). Fig. 2. Experimental assembly.
E. Croner et al. / International Journal of Heat and Fluid Flow 43 (2013) 233–243 235

used to obtain mean and instantaneous information on velocity


fields in the wake of an isolated wheel rotating in contact with
the ground (see Fig. 2).
The ACE wind tunnel featured a 2.35 m-wide and 2 m-high test
section. Hence it was possible to test a full scale wheel assembly
which was composed of a strut and a wheel. The strut can be
divided into two parts, one mast and one arm, which were linked
together by a pivot joint. Both parts were given streamlined
cross-sections so as to be as unobtrusive as possible. The wheel
was carefully aligned with the oncoming flow and only a 0° yaw
angle is studied here. In order to simulate real road conditions,
the test section was equipped with a moving belt to provide both
ground simulation and rotation. The belt velocity was synchro-
nized with the air free-stream, and the tests were run at the se-
lected value of 22 m/s, corresponding to a Reynolds number
based on the wheel diameter of 9.1  105. A suction system was
placed ahead of the belt in order to remove the boundary layer
developing on the lower wall of the convergent.
The tyre complied with the 205/55/R16 standard thus corre-
sponding to a width of L = 0.205 m and an external diameter of
D = 0.631 m. Both sides of the rim were covered with smooth black
carbon plates. The smooth shaped tyre was specially designed for
this experiment by the French tyre manufacturer Michelin in order
to exhibit a flat tread when rotating. Indeed, a low load level was
required so as not to damage the belt. The flat part provides a con-
tact along the full width of the tread even with low load. Two
adjusting screws enabled the position of the arm to be set so as
to find the configuration that provided the minimal load for which
there was no slippage between the tyre and the moving belt. As a
result, the deformation of the tyre was minimal but the contact
width was realistic.
A 2D-2C PIV system is used to record the flow velocity field
information. Illumination is provided by a double pulsed Nd:YAG
laser emitting two pulses of 120 mJ each (laser sheet thickness
1 mm). An oil generator provides seeding with a mean particle
diameter of about 1 lm. Two LaVision CCD cameras, both with a
1376  1040 pixel square resolution were used. The pre-process-
ing corresponds to the subtraction of the background image using
the minimum value within the set of images. Velocity vector fields
Fig. 3. PIV planes: constant X (a), constant Y and constant Z (b).
are computed with LaVision 7.2 software. A multipass algorithm
with a final interrogation window size of 16  16 pixels and 50%
overlapping is applied. For each position in Fig. 3, 2000 image pairs and the Z axis is vertical. The test section of the wind tunnel is
were obtained at a frequency of 4 Hz. Consequently, estimated sta- reproduced with the exact moving belt and wheel positions (see
tistical absolute errors for mean and root-mean-square (RMS) val- Fig. 4); only its length has been increased at each extremity to min-
ues (Benedict and Gould, 1996) are respectively DU  0.04uRMS and imize the influence of the boundary conditions. The suction system
DuRMS  0.03uRMS with a 95% confidence level. is replaced by an upstream extension of the moving belt in order to
Further investigations were performed in the wake of the rotat- avoid the boundary layer to develop.
ing wheel using a one-dimensional hot-wire anemometer. The Calculations are based on an accurate representation of the
sampling frequency was set at 5 kHz with a duration of two min- smooth rotating wheel used for experiments and a simplified ver-
utes for each measurement. The time series data are split into seg- sion of the strut without links between the mast and the arm. The
ments of 8092 points with a 50% overlap and a Hamming window low load level of the wheel used for experiments minimizes the
is used to compute the modified periodogram of each segment. The tyre deformation and its flat tread fixes its width. Hence, the tyre
purpose of this study was to probe locations that featured vortical geometry can be approximately obtained by revolution of the real
structures according to previous PIV measurements. In the plane profile of the tyre taking the centrifugal force into account. The
X = D downstream from the wheel centre, velocities were recorded ground contact is produced by cutting the wheel at an appropriate
at different heights above the ground in the median plane and ground level (0.5 mm), thus obtaining the correct dimensions for
downstream of the jetting-vortex on the strut-free side. the contact patch.
Inflow conditions are defined to match experimental ones with
a uniform velocity profile at Vi = 22 m/s which gives a Mach num-
3. Numerical modelling ber of 0.064 and a Reynolds number based on the wheel diameter
of 9.1  105. The flow can be considered as fully turbulent. A rela-
CFD simulations of the experimental configuration were per- tive turbulence intensity of 0.1% and an eddy viscosity ratio of 0.1
formed in order to provide a direct comparison with experimental are defined to calculate turbulence variables at the inlet. At the
results. The origin of the cartesian coordinate system is placed at outlet, constant static pressure is prescribed. A slip boundary con-
the centre of the wheel. The X axis is parallel to the moving floor dition is specified on the wind tunnel walls to avoid resolving the
and oriented downstream, the wheel rotates around the Y axis boundary layers along them. The floor, the moving belt, the wheel
236 E. Croner et al. / International Journal of Heat and Fluid Flow 43 (2013) 233–243

found to be the most closest at predicting the time and space char-
acteristics of the flow in this configuration.
The spatial discretization uses a second-order accurate upwind
scheme and the global time-step used for the explicit time resolu-
tion is 105 s, ensuring a Courant–Friedrichs–Lewy number CFL 6 1
everywhere. This time step is inferior to the required time-step in
the wake to converge but it is fixed to minimize the CFL number
around the contact area, where important pressure peaks occur.
As a result, the CFL number in the wake is less than 0.1. For com-
parison, Axerio-Cilies et al. (2012) found a time-step of 104 for
reaching time-step independency. Finally, the Weiss–Smith low-
Mach preconditioning (Weiss and Smith, 1995) is applied.
It must be mentioned that, in the case of an isolated rolling
wheel in contact with a moving ground, steady RANS solutions
cannot converge unless adding numerical dissipation by using
coarse grids, artificial viscosity or dissipative integration schemes.
This issue, already mentioned in previous studies (Basara et al.,
2000; Axerio-Cilies et al., 2012), is due to the intrinsic unsteadiness
of the wake of bluff bodies, linked to vortex shedding phenomena.
The difficulty to get steady solutions was also observed for isolated
fixed wheels (Axerio et al., 2009) and rotating wheels in wheel-
houses (Axon et al., 1999).

4. Results

4.1. Pressure distribution

The distribution of averaged pressure coefficient CP along the


wheel centre-line in Fig. 5 shows that the characteristic features
mentioned in the litterature are reproduced by URANS calcula-
tions. For h 6 90°, the pressure differences outline higher velocities
caused by the lower aspect ratio of the present geometry. The po-
sitive and negative pressure peaks CPmin = 13.5 and CPmax = 15.6
around the contact position (h = 90°) characterize Fackrell’s viscous
pumping phenomenon. The 100–270° section (base part) is found
to be strongly unsteady, confirming the observations of Fackrell
and Harvey (1974). The discrepancies in this area are linked to
the recirculation area whose aerodynamic characteristics depend
on the Reynolds and the wheel geometry (for example aspect ratio
and tyre shoulders). This section ends by a depression just up-
stream of the top of the wheel (277°), referred as suction peak.
Fig. 4. Computational domain (a) and meshing of the walls (b).

4.2. Upper boundary layer separation


and the strut surfaces are considered as adiabatic walls with no-
The location of the boundary layer separation on the top of a
slip condition. The moving belt and the wheel velocities are ad-
rotating wheel is one of the characteristics used for the validation
justed to match the inflow velocity Vi, with an adequate angular
of numerical results. An inaccurate prediction of this separation
velocity of 69.7 rad/s for the wheel.
The block-structured computational mesh (see Fig. 4b) was cre-
ated with ICEM CFD software. It consists of 9  106 hexahedral
cells distributed on 4 non-matching subdomains: the main domain
around the wheel and 3 smaller ones fitting the support. The min-
imal cell size at the walls enables a correct resolution of the bound-
ary layers according to wall unit values (y+ 6 1). The density of
cells in the wake and near-wall regions is sufficient to obtain grid
independent solutions and comparable to the one of Axerio-Cilies
et al. (2012).
The ONERA code elsA (Cambier and Gazaix, 2002) has been used
to solve URANS equations by the finite volume discretization
method. The ONERA k–kL two-equation turbulence model (Bézard
and Daris, 2005) was chosen for turbulence closure. The aim of the
present paper is not to present a study of the numerical parameters
and their accuracy in the case of the rotating wheel. Nonetheless
two other eddy viscosity models used in the industry, the k–x
SST (Menter, 1994) and Spalart–Allmaras (Spalart and Allmaras,
1992) models, have been tested and the k–kL model has been Fig. 5. CP distribution along the wheel centre-line.
E. Croner et al. / International Journal of Heat and Fluid Flow 43 (2013) 233–243 237

and of the recirculation area just downstream can lead to major


discrepancies with experiments concerning the velocity levels
and the flow topology (Axerio-Cilies et al., 2012).
The location of this separation is generally deduced from the CP
distribution around the centre-line. Fackrell and Harvey (1974)
mention that the averaging of the base pressure gives a region of
quasi-constant CP and that the end of this region is interpreted as
the angular position of the separation. However the precision of
this criterion is open to discussion. A suction peak often occurs
in this area whose amplitude depends on several criteria, including
for example the tyre profile (Fackrell and Harvey, 1974; Skea et al.,
Fig. 7. Top view of iso-contours of friction (solid) and pressure coefficient (dashed)
2000). This peak is also very sensitive to the turbulence models on the wheel surface.
(Basara et al., 2000; Régert and Lajos, 2002). As a consequence
the end of the base pressure section is not clearly defined. Most (MRS) criteria, based on a zero radial acceleration of the tangential
authors consider the separation to occur downstream of the suc- velocity, to detect the saddle point.
tion peak.
However, to the authors’ knowledge, the link between CP distri-
bution and upper separation is not established for a rotating wheel. 4.3. Analysis of the flow topology
The saddle point topology first suggested by Fackrell and Harvey
(1974) was confirmed by theoretical and experimental investiga- Numerical results were time-averaged over 2  105 time-steps
tions (Régert and Lajos, 2002) as well as numerical simulations and compared to PIV results (see Fig. 8). Statistical studies at differ-
(Pirozzoli et al., 2012). But low-Reynolds DNS of Pirozzoli et al. ent positions in the wake confirmed the convergence of mean and
(2012) detect the saddle point at h = 285° while the CP distribution RMS values after 1  105 time steps. The boundary layer separation
exhibits a continuous increase in pressure after 270°. induces a recirculation bubble predicted further downstream by
In this study, the authors found that the saddle point character- CFD compared to PIV. The position of the separation and the bub-
izing this separation does not match any specific position of the CP ble shape are quasi-steady in both experiments and numerical sim-
distribution. In the numerical results, this saddle point is found at ulations. Streamlines plotted in the median plane highlight the
hSP = 291° while the local minimum of CP is located at hC P ¼277°. A downwash phenomenon: the flow coming from above the wheel
visualization of this saddle point can be obtained by a rotation of is driven downwards by the rotation of the vortices.
the flow field of hSP and a dilatation of the normal ZROT axis as The numerical model used in this paper over-estimates the
shown in Fig. 6. velocity in the upper part of the wake and the intensity of the recir-
The analysis of different variables around these two particular culation area in the lower part with lower negative values near the
positions highlights the occurrence of two different phenomena. ground. This can be seen by the comparison of streamwise velocity
The saddle point occurs 3° downstream from a critical point in profiles in the midplane, plotted in Fig. 9). However the global flow
the wall friction field: a local maximum in the streamwise direc- topology is well reproduced considering the number and the posi-
tion X and a local minimum in the spanwise direction Y (Fig. 7). tion of vortices. This difficulty to precisely predict the velocity lev-
It also matches a local maximum of kinetic turbulent energy in els in the downwash area with 2-equation turbulence models was
the boundary layer and the first angular position for which a posi- also observed by Axerio-Cilies et al. (2012), with errors up to
tive normal velocity appears on the top of the wheel. ±10 m/s in this area for an upstream velocity of 18.4 m/s.
The suction peak, characterized by a local minimum of CP lo- The wake dimensions can be observed for different X thanks to
cated 14° upstream of the saddle point, seems to be linked to an- in-plane velocity (VP) contours in Fig. 10. This can be completed by
other phenomenon: a local acceleration just above the boundary the C2 criterion (Graftieaux et al., 2011) and Q criterion (Hunt
layer. et al., 1988) which detect coherent vortical structures. The C2 cri-
It must be mentioned that the extraction of velocity profiles terion is based on the flow topology in a 2D area and is therefore
around the wheel does not show any vertical tangent which is in suitable for analyzing two-components PIV measurements. The Q
contradiction with the suggestions of Fackrell and Harvey (1974) criterion is calculated from all components of the velocity gradient
and those of Régert and Lajos (2002) about the saddle point char- by Q ¼ 12 ðkXk2  kSk2 Þ, where X is the vorticity tensor and S the
acteristics. Hence it was not possible to use the Moore–Rott–Sears strain-rate tensor. The local flow is considered to be dominated
by the rotation when Q  0 and jC2j > 2/p. C2 criterion contours
extracted from PIV and URANS simulations are compared in
Fig. 11. The Q criterion is also plotted in the same plane for numer-
ical results in Fig. 12. These results, added to the 3D iso-surface of Q
criterion in Fig. 13, exhibit massive coherent structures and enable
visualization of the wake development.
In the near-wake, namely X 6 D, both PIV and CFD results show
the upper vortices 1 and 2 (see Fig. 12) caused by the boundary
layer separation and the near-ground vortices 3 and 4 linked to
the jetting, as well as an additional pair of vortical structures (5
and 6) as suggested by Pirozzoli et al. (2012). Vortices 5 and 6, orig-
inating from the wheel base, merge either with one upper vortex or
with a lower one, resulting in vortex size differences at X = D. The
weak vortices that have been suggested on centres of the sides of
the wheel are not observed. In addition, as observed by Wäschle
(2007), the two lower vortices do not originate directly from the
jetting vortices but are actually connected in a massive wake-
Fig. 6. Visualisation of the saddle point in CFD results through streamlines. horseshoe vortex like the one isolated in Fig. 14.
238 E. Croner et al. / International Journal of Heat and Fluid Flow 43 (2013) 233–243

(a) CFD (b) PIV


Fig. 8. Streamlines and VX contours in the midplane Y = 0.

Fig. 9. Profiles of VX in the midplane Y = 0.

The correlations between the four main vortices have been cal- hðCi  hCi iÞðCj  hCj iÞi
Rij ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð1Þ
culated from experimental results thanks to their circulation C at
hðCi  hCi iÞ2 ihðCj  hCj iÞ2 i
X ¼ 34 D. The correlation coefficient is defined in Eq. (1), where the
average operator is noted hi. The upper vortices 1 and 2 have been
found to be better correlated one to each other compared to lower
vortices 3 and 4 (see Table 1). Moreover the behaviour of upper 4.4. Analysis of the flow unsteadiness
vortices does not seem to be linked to that of the lower ones.
The circulation density function, not presented in this paper, also 4.4.1. Velocity spectra
highlights complex behaviour with some changes in the direction The flow unsteadiness is first studied at different positions
of rotation for lower vortices. in the wake through velocity evolutions given by numerical
E. Croner et al. / International Journal of Heat and Fluid Flow 43 (2013) 233–243 239

(a) CFD (b) PIV


Fig. 10. VP contours in the plane X = D.

(a) CFD (b) Experiments


Fig. 11. Comparison of numerical (a) and experimental (b) C2 contours in the plane X = 0.66D.

calculations and hot-wire anemometry. Velocity spectra are ob-


tained by discrete Fast Fourier Transformations (FFT). Some results
are presented in Figs. 15 and 16 through Power Spectral Density
(PSD) plotting. For each component and each position in the wake,
the main frequencies detected by hot-wire anemometry can be di-
vided into two parts: 11.1 Hz and its harmonics on the one hand,
14.8 Hz and its harmonics on the other. It has been found that
the first frequency matches the wheel rotation frequency while
the wake dynamics are characterized by 14.8 Hz and its first har-
monic, as explained in Section 5. The use of the URANS method
provides numerical spectra with strong peaks corresponding to
the main motions of the wake. The main frequencies are 12.5 Hz
and its harmonics which can be compared to 14.8 Hz. As a result,
the temporal characteristics of the numerical wake show good con-
cordance with experiments considering the use of URANS simula-
tions. The Strouhal numbers based on the wheel width
corresponding to the main frequency are StW exp = 0.137 for the
Fig. 12. Numerical contours of Q criterion in the plane X = 0.66D. experiments and StW num = 0.116 for the numerical simulations.
240 E. Croner et al. / International Journal of Heat and Fluid Flow 43 (2013) 233–243

Fig. 13. Iso-surface Q = 1000.


Fig. 15. Velocity spectra in the plane X = D for experimental results.

Fig. 14. Bottom view of some lower-wake vortical structures (Q = 10,000).

Table 1 Fig. 16. Velocity spectra in the plane X = D for numerical results.
Correlation coefficients of vortex circulations.

R12 R34 R13 R24


0.554 0.075 0.053 0.162
Table 2
Aerodynamic coefficients of the wheel extracted from URANS calculations.

CD CL CS
However the Strouhal number is not precisely defined in the case
of a rotating wheel (see Section 5). Mean value 0.444 0.164 0.0331
Variations ±0.0125 ±0.0145 ±0.0775
The main frequency is now referred as f1 and its first harmonic
f2. In both hot-wire and CFD results, f1 is predominant in velocity
variations for all probe positions in the wake. Time variations are contact between the tyre and the moving belt. CFD simulations
particularly significant downstream of the jetting-vortices near give access to the mean values and time evolutions of the aerody-
the ground with large Vx and Vy variations. In the median plane, namic force coefficients, presented in Table 2 and Fig. 17, where CD,
Vy variations remain high while Vx variations decrease. For numer- CL and CS refer respectively to drag, lift and side force coefficients.
ical simulations, additional probes were placed just behind the jet- Spectral analysis of these results reveals the same characteristic
ting origin and the upper separation in the midplane. The upper frequencies f1 and f2 as previously.
separation stability is confirmed, with velocity variations below The side force coefficient exhibits the largest variations with a
1%, and the area just behind the contact zone is found to be dom- dominant frequency equal to f1 = 12.5 Hz. Most of time, this force
inated by f2. is negative because of the set-up asymmetry, which also induces
a global deviation of the wake in the direction of Y P 0. Consider-
4.4.2. Aerodynamic coefficients ing this orientation, the wheel behaves just like an asymmetric
The aerodynamic forces acting on the wheel were not measured wing profile producing a negative mean side force. Similarly to
during the test campaign. Such measurements are difficult to per- the signal in the jetting area, drag and lift coefficients exhibit
form and require to overcome the components linked to the slightly higher spectral density for f2 = 25 Hz.
E. Croner et al. / International Journal of Heat and Fluid Flow 43 (2013) 233–243 241

(a) 0 s – CS max.

Fig. 17. Temporal evolution of aerodynamic coefficients over 2 s (2  105 time-


steps) extracted from URANS calculations.

Moreover, the viscous components of each force were found to


be quasi-steady; consequently, their variations are due solely to
pressure modifications.

4.4.3. Vortical structures


The temporal analysis of an iso-Q criterion surface over one per-
iod of 0.08 s, matching f1 = 12.5 Hz, is plotted in Fig. 18, where cap- (b) 0.02 s
ital letters indicate the dominant structure between the left and
right sides. It reveals how the wake 3D vortical structures are cre-
ated and how the above-mentioned wake-horseshoe vortex is clo-
sely related to the jetting phenomenon.
At t = 0 s, the structures A and a are linked in a wake-horseshoe
vortex, attached to the wheel base-surface, while the jetting phe-
nomenon produces B and b. At the following instant, B and b start
interacting with the highly unsteady wheel base area. When these
structures are shed in the wake, at t = 0.04 s, they are linked in a
new wake-horseshoe vortex and the predominance of B is empha-
sized. In the meantime, C and c are produced by the jetting. As a
result, the massive 3D structures in the wake are created by the
interaction between the quasi-symmetric structures originating
from the jetting and the unsteady wheel base flow. This issue is
(c) 0.04s – CS min.
discussed in Section 5.

5. Discussion

Contour plotting shows good concordance between numerical


and PIV results. Despite some areas with higher velocities, the
structures developed in the wake are well reproduced. CFD results
also reproduce the characteristic flow features mentioned in the
literature such as the jetting phenomenon. Spectral analysis cou-
pled to the above-mentioned results on the flow topology makes
it possible to validate the numerical model before carrying out a
detailed analysis of the wake dynamics.
The ring vortex suggested by Wäschle (2007) on the upper-part
of the wheel has not been observed. The results show an arch-
shaped vortex (I in Fig. 19) just downstream from the top of the (d) 0.06 s
wheel, matching the recirculation bubble in Fig. 8, associated with
the highly correlated contra-rotative vortex pair (II in Fig. 19). Fig. 18. Top views of iso-surface Q = 5000 at 4 instants of the 0.08 s period.
These steady structures are produced by the upper wheel bound-
ary layer separation and the wheel rotation. The vortex pair be- with the unsteady behaviour of the depression area downstream
comes unsteady only after the downwash when it interacts with of the contact between the wheel and the moving belt. On both
the lower part of the wheel wake. sides of the contact zone, the jetting vortices are quasi-symmetric
URANS calculation reveals complex structures for the lower and structures are simultaneously shed at f2 frequency. When they
part of the wake. The jetting phenomenon appears to be coupled interact with the flow behind the wheel, the jetting vortices lose
242 E. Croner et al. / International Journal of Heat and Fluid Flow 43 (2013) 233–243

therefore essentially controlled by f2 frequency giving numerical


Strouhal numbers of StW = 0.233, based on the wheel width, and
StD = 0.717, based on the wheel diameter. To the authors knowl-
edge, there are few evaluations of the Strouhal number in the liter-
ature and only for wheels in wheelhouses. URANS results of
Damiani et al. (2004) give StW = 0.22 and Krajnović et al. (2011)
find a dominant peak at StD = 1.39 for the case of a wide wheel-
house. However these studies were carried out at a subcritical Rey-
nolds number (Cogotti, 1983 gives ReC = 106) and it is difficult to
compare isolated wheel results with those of vehicles.
Finally, the configuration of a rotating wheel contacting the
ground also leads to some definition problems linked to the flow
specificities.
The rotation of the wheel causes the upper separation to be
characterized by a saddle point topology just above the wall as de-
scribed in Section 4.2. This separation is not visible by friction line
plotting and therefore needs other criteria to be detected. For
numerical simulations, the authors suggest using local variations
of friction instead of the variation of CP or MRS criterion, although
it is possible that the separation also has an influence on the CP dis-
Fig. 19. Iso-surface Q = 1000 and streamlines. tribution. Further study about the separation and its link with the
suction peak may be carried out.
In addition, the characteristic length is not clearly defined for
their symmetry because of the base flow lateral motion, one struc- the analysis of the frequential behaviour. The diameter is adapted
ture becoming predominant. to the case of infinite cylinder but in the case of wheels the wake is
For the configuration under consideration, the asymmetry is in- necessarily affected by the wheel width. Hence, the authors sug-
creased by that of the global flow caused by the set up. A horseshoe gest that the characteristic length of a wheel (low aspect ratio cyl-
vortex is then created from the interaction of the two structures inder) may be a combination of diameter and width.
with the wheel rotation in the base recirculation area. The struc-
tures observed in the wake are those shed from this massive horse- 6. Conclusion
shoe vortex. For each vortex pair, the weaker side is dissipated far
more quickly as seen in Fig. 18d with the pairs A–a and B–b. The Wind tunnel testing and numerical modelling were carried out to
main frequency in the wake is consequently reduced to f1. characterize the wake of an isolated wheel. The CFD model has been
Hence the wake does not exhibit a classical vortex shedding able to reproduce both the flow topology and its unsteady behaviour
phenomenon similar to those producing the von Kármán vortex with correct characteristic frequencies. As a result, the analysis of
street. The structures originating from upstream of the contact area the vortices dynamics and the identification of their effect on the
are affected by the downstream area and become asymmetric. The aerodynamic coefficients give new insight into the unsteady evolu-
weak correlation coefficients of the lower vortical structures are tion of the wake and the flow surrounding the wheel. It facilitates the
probably linked to this complex unsteady phenomenon which understanding of the mechanisms creating the main vortices in the
causes this structures to be less stable than the upper ones. wake of the isolated wheel and of their interaction.
The use of CFD enables further understanding of the flow by These observations may be interesting for future studies in or-
analyzing the effects of vortex dynamics on the aerodynamic coef- der to:
ficients evolutions. The three coefficient time-evolutions are con-
trolled by the lower part of the wake. The drag and lift  understand the influence of the wheel geometry and the wheel-
coefficients are equally affected by f1 and f2 frequencies while the house on the vortical structures;
side force coefficient evolution exhibits only one characteristic fre-  understand the interaction between a vehicle body and the
quency f1. wakes of the wheels;
CD and CL evolutions appears to be strongly correlated – maxi-  elaborate new solutions for reducing the aerodynamic drag of
mal drag matching minimal lift – and controlled by the vortex vehicles.
pairs shedding frequency f2. This behaviour is closely related to
the nearwake lower part where a pressure decrease induces lower
lift and higher drag. Hence, it has been found that maximal drag Acknowledgements
matches the moment when the jetting vortices interact with the
wake-horseshoe vortex and make it become larger (see Fig. 18b The authors thank B. Gardarin, the project manager at Michelin;
and d). This phenomenon can be compared to the observations of J. Borée and L.-E. Brizzi of Institut Pprime for having conducted the
Krajnović et al. (2011) in the case of a wheel included in a wheel- wind tunnel experiments; P. Braud of Institut Pprime for his tech-
house: Large Eddy Simulation (LES) results indicated that an in- nical assistance throughout the test campaign; and the Aero Con-
crease in density of flow structures in the wheel wake is visible cept Engineering society. The research was supported by Michelin.
when the drag is maximum. In contrast, CS is affected by pressure
distributions on the sides of the wheel and therefore highly de-
References
pends on the predominance of one jetting vortical structure. The
link between the wake and the side force can be illustrated in Axerio-Cilies, J., Issakhanian, E., Jimenez, J., Iaccarino, G., 2012. An aerodynamic
Fig. 18 where (a) occurs for the maximum value of CS and (c) for investigation of an isolated stationary formula 1 wheel assembly. Journal of
the minimal one. The authors suggest that the predominance of Fluids Engineering 134 (021101).
Axerio, J., Iaccarino, G., Issakhanian, E., Lo, K., Elkins, C., Eaton, J., 2009.
one of the two structures is less significant in the case of an iso- Computational and Experimental Investigation of the flow Structure and
lated wheel without strut. The drag and lift evolutions are Vortex Dynamics in the Wake of a Formula 1 Tire. SAE Paper (2009-01-0775).
E. Croner et al. / International Journal of Heat and Fluid Flow 43 (2013) 233–243 243

Axon, L., Garry, K., Howell, J., 1999. The Influence of Ground Condition on the Flow McManus, J., Zhang, X., 2006. A computational study of the flow around an isolated
Around a Wheel Located Within a Wheelhouse Cavity. SAE Paper (1999-01- wheel in contact with the ground. Journal of Fluids Engineering 128 (3), 520–
0806). 530.
Basara, B., Beader, D., Przulj, V., 2000. Numerical simulation of the air flow around a Mears, A., Dominy, R., Sims-Williams, D.B., 2002. The Airflow About an Exposed
rotating wheel. In: 3rd MIRA International Vehicle Aerodynamics Conference, Racing Wheel. SAE Paper (2002-01-3290).
Rugby, UK. Menter, F.R., 1994. Two-equation eddy-viscosity turbulence models for engineering
Benedict, L., Gould, R., 1996. Towards better uncertainty estimates for turbulence applications. AIAA Journal 32 (8), 1598–1605.
statistics. Experiments in Fluid 22 (2), 129–136. Mercker, E., Berneburg, H., 1992. On the simulation of road driving of a passenger
Bézard, H., Daris, T., 2005, Calibrating the length-scale equation with an explicit car in a wind tunnel using a moving belt and rotating wheels. In: 3rd
algebraic Reynolds stress model. In: ERCOFTAC International Symposium on International Conference Innovation and Reliability, Florence, Italy.
Engineering Turbulence Modelling and Experiments-ETMM6, Sardinia, Italy. Mlinaric, P., Sebben, S., 2008. Investigation of the influence of tyre deflection and
Cambier, L., Gazaix, M., 2002. elsA: An efficient object-oriented solution to CFD tyre contact patch on cfd predictions of aerodynamic forces on a passenger car.
complexity. In: 40th AIAA Aerospace Sciences Meeting and Exhibit. In: 7th MIRA International Vehicle Conference, PN 116-022.
Cogotti, A., 1983. Aerodynamics characteristics of car wheels. International Journal Modlinger, F., Demuth, R., Adams, N., 2007. Investigations on the Realistic Modelling
of Vehicle Design, 173–196 (Special Publication SP3). of the Flow Around Wheels and Wheel Arches by CFD. JSAE Paper (20075195).
Damiani, F., Iaccarino, G., Kalitzin, G., Khalighi, B., 2004. Unsteady flow simulation of Pirozzoli, S., Orlandi, P., Bernardini, M., 2012. The fluid dynamics of rolling wheels at
wheel-wheelhouse configurations. In: 34th AIAA Fluid Dynamics Conference low Reynolds number. Journal of Fluid Mechanics 706.
and Exhibit, Portland, USA. Régert, T., Lajos, T., 2002. Numerical simulation of flow field past road vehicle
Elofsson, P., Bannister, M., 2002. Drag Reduction Mechanisms due to Moving wheel. In: Proceedings of Gépészet 2002 Conference, pp. 244–248.
Ground and Wheel Rotation in Passenger Cars. SAE Paper (2002-01-0531). Régert, T., Lajos, T., 2007. Description of flow field in the wheelhouses of cars.
Fackrell, J., Harvey, J., 1974. The aerodynamics of an isolated wheel. In: Proceedings International Journal of Heat and Fluid Flow 28 (4), 616–629.
of the 2nd AIAA Symposium on the Aerodynamics of Sport and Competition Saddington, A., Knowles, R., Knowles, K., 2007. Laser Doppler anemometry
Automobiles, vol. 16, pp. 119–125. measurements in the near-wake of an isolated formula one wheel.
Graftieaux, L., Michard, M., Grosjean, N., 2011. Combining PIV, POD and vortex Experiments in Fluid 42 (5), 671–681.
identification algorithms for the study of unsteady turbulent swirling flows. Skea, A., Bullen, P., Qiao, J., 2000. CFD Simulations and Experimental Measurements
Measurement Science and Technology (12), 1422–1429. of the Flow Over a Rotating Wheel in a Wheel Arch. SAE Paper (2000-01-0487).
Hunt, J., Wray, A., Moin, P., 1988. Eddies, Stream, and Convergence Zones in Spalart, P.R., Allmaras, S.R., 1992. A one-equation turbulence model for
Turbulent Flows. Technical Report. Center for Turbulence Research. aerodynamic flows. In: AIAA 30th Aerospace Sciences Meeting And Exhibit
Iaccarino, G., Durbin, P., 2000. Unsteady 3D RANS simulations using the v2-f model. (AIAA 92-439).
In: Annual Research Briefs. Center for Turbulence Research, Stanford University, Wäschle, A., 2007. The Influence of Rotating Wheels on Vehicle Aerodynamics-
pp. 263–269. Numerical and Experimental Investigations. SAE Paper (2007-01-0107).
Krajnović, S., Sarmast, S., Basara, B., 2011. Numerical investigation of the flow Weiss, J., Smith, W., 1995. Preconditioning applied to variable and constant density
around a simplified wheel in a wheelhouse. Journal of Fluids Engineering flows. American Institute of Aeronautics and Astronautics Journal 33 (11),
(111001). 2050–2057.

You might also like