25 Degree Ahmed Body aerodynamic study

Embed Size (px)

Citation preview

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    1/19

    13

    Experiments in Fluids

    Experimental Methods and their

    Applications to Fluid Flow

    ISSN 0723-4864

    Volume 52

    Number 5

    Exp Fluids (2012) 52:1169-1185

    DOI 10.1007/s00348-011-1245-5

    Drag reduction on the 25 slant anglehmed reference body using pulsed jets

    Pierric Joseph, Xavier Amandolse &

    Jean-Luc Aider

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    2/19

    13

    Your article is protected by copyright and

    all rights are held exclusively by Springer-

    Verlag. This e-offprint is for personal use only

    and shall not be self-archived in electronic

    repositories. If you wish to self-archive yourwork, please use the accepted authors

    version for posting to your own website or

    your institutions repository. You may further

    deposit the accepted authors version on a

    funders repository at a funders request,

    provided it is not made publicly available until

    12 months after publication.

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    3/19

    RE S E A RCH A RT I CL E

    Drag reduction on the 25

    slant angle Ahmed referencebody using pulsed jets

    Pierric Joseph Xavier Amandolese

    Jean-Luc Aider

    Received: 11 June 2011 / Revised: 13 November 2011 / Accepted: 29 November 2011 / Published online: 15 December 2011

    Springer-Verlag 2011

    Abstract This paper highlights steady and unsteady

    measurements and flow control results obtained on anAhmed model with slant angle of 25 in wind tunnel. On

    this high-drag configuration characterized by a large sep-

    aration bubble along with energetic streamwise vortices,

    time-averaged and time-dependent results without control

    are first presented. The influence of rear-end periodic

    forcing on the drag coefficient is then investigated using

    electrically operated magnetic valves in an open-loop

    control scheme. Four distinct configurations of flow control

    have been tested: rectangular pulsed jets aligned with the

    spanwise direction or in winglets configuration on the roof

    end and rectangular jets or a large open slot at the top of the

    rear slant. For each configuration, the influence of the

    forcing parameters (non-dimensional frequency, injected

    momentum) on the drag coefficient has been studied, along

    with their impact on the static pressure on both the rear

    slant and vertical base of the model. Depending on the type

    and location of pulsed jets actuation, the maximum drag

    reduction is obtained for increasing injected momentum or

    well-defined optimal pulsation frequencies.

    1 Introduction

    Current environmental and economic issues lead automo-

    tive manufacturers to search for innovative solutions to

    reduce vehicles fuel consumption. One way is to reduce

    aerodynamic drag, which is responsible for the largest

    part of the fuel consumption for speed above 80 km h-1

    (Hucho1998). Like bluff-body, automotive drag is mainly

    governed by massive separation on the rear part: for a

    typical family car, pressure drag on this area can reach near

    a third of the total aerodynamic drag (Barnard 1996).

    In order to simplify the study of automotive near wake,

    and thus to understand aerodynamic drag generation in the

    rear part of a vehicle, Ahmed et al. (1984) introduced a

    simplified geometry (Fig.1a). Despite the fact that this

    geometry is close to its thirtieth anniversary, it is still lar-

    gely used by scientific community as an automotive ref-

    erence model to work on complex three-dimensional wake

    flow and its control, using numerical methods (Krajnovic

    and Davidson 2005a, b; Fares 2006) and experimental

    techniques (Beaudoin et al.2004; Thacker2010; Gillieron

    2010).

    The flow topology of the Ahmed body, and thus its

    aerodynamic drag, is greatly dependent of the slant angle.

    As this angle evolves from 0 to 90, the near wake of the

    Ahmed bluff-body changes drastically. From 12 to 15,

    the flow is typical of a rear blunt with a flow separation on

    rear edges, generating mainly transverse vorticity: this first

    type of separation is sometimes considered as quasi-two-

    dimensional (Hucho 1998). From 15 up to 30, the near

    wake is highly three-dimensional with partial separation on

    the slant surface along with strong conical streamwise

    vortices coming from the slant side edges and a ring-

    shaped structure lying on the base surface. Beyond 30, the

    separation can also be considered as quasi-two-dimensional

    P. Joseph

    Institut AeroTechnique (IAT), CNAM, 15 rue Marat,78210 Saint Cyr lEcole, France

    e-mail: [email protected]

    X. Amandolese

    Aerodynamics Department, CNAM, 15 rue Marat,

    78210 Saint Cyr lEcole, France

    e-mail: [email protected]

    J.-L. Aider (&)

    PMMH Laboratory, UMR 7636, CNRS, ESPCI ParisTech,

    10 rue Vauquelin, 75231 Paris, France

    e-mail: [email protected]

    1 3

    Exp Fluids (2012) 52:11691185

    DOI 10.1007/s00348-011-1245-5

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    4/19

    due to the massive separation from the top of the rear slant.

    Figure1b is a schematic view of the mean flow topology

    for the three-dimensional situation (Vino et al. 2005).

    This three-dimensional complex wake exists with a 25

    slant angle and presents high-drag coefficient, which makes

    it a good test case for drag reduction study.

    After years of drag reduction using shape optimization,

    this technique shows its limit regarding design constraints

    of the automotive industry. This trend causes flow control

    techniques to be more and more studied, with the Ahmed

    body as a benchmark.

    A lot of successful studies can be found in the literatureusing passive strategies: for example, Fourrieet al. (2011)

    obtained 9% of drag reduction using a classical automotive

    style deflector, while Beaudoin and Aider (2008) reached

    an impressive 25% reduction with several flaps located on

    the edges of the rear end of a 30 configuration. An

    approach using vortex generators to produce coherent

    streaks that increase or decrease the separation bubble was

    also carried out by Aider et al. (2009) and Pujals et al.

    (2010) leading, respectively, to a 12 and 10% drag reduc-

    tion. However, apart from the active vortex generators

    proposed by Aider et al. (2009), all these passive tech-

    niques introduce quite unsightly appendages on the body,which is in contradiction with design constraints.

    This fact makes active control by jet or suction very

    attractive in automotive industry. This kind of control is

    nearly invisible and can be adapted to changes in flow

    conditions. Various successful active control studies have

    been conducted using the Ahmed reference body with 25

    slant angle. Roumeas et al. (2008) used steady aspiration on

    the top of the slant: he obtained numerically a drag reduc-

    tion of 17% and noticed experimentally a suppression of the

    separation area (Roumeas2006). Experimental studies were

    carried out by Leclerc (2008) with synthetic jets (zero net

    mass flux) at the top slant edge area (8.5% reduction) and by

    Krentel et al. (2010) with pulsed jets at the bottom slant

    edge (5.7% reduction). Both Krajnovic et al. (2009) and

    Lehugeur (2009) made numerical simulations of the same

    case: the former obtained a little bit more than 7% reduction

    using steady blowing and suction at the slant top edge (and

    also studied several other blowing locations and jet types),

    while the latter used steady blowing to force the bursting of

    longitudinal coherent structures, leading to a 6% drag

    reduction. Brunn et al. (2008) have experimented anadvanced control strategy by targeting simultaneously

    particular structures with different actuator types (steady

    and periodic), including closed-loop features. Beaudoin

    et al. (2008) used as well feedback control by extremum

    seeking on a rounded Ahmed Body. Periodic forcing was

    also successfully used by Pastoor et al. (2008) on a more

    simplified body. They managed to reduce drag by 15% by

    synchronizing upper and lower vortex shedding with a

    synthetic jet control system in closed loop.

    According to those studies, it seems that periodic forcing is

    a promising way of controlling the flow structures on the 25

    Ahmed body. A good understanding of this kind of control isachieved on academic geometry, like backward-facing step

    (see Tihon et al. 2010). Optimal frequencies are clearly

    identified among the base flow instabilities, as theeffect of the

    forcing amplitude. Meanwhile, on three-dimensional com-

    plex flow like theAhmed body wake, influenceof each forcing

    parameters (injected momentum, non-dimensionalfrequency,

    spanwise modulation, etc.) is still under discussion.

    In the present paper, we will focus on the suppression

    of the rear slant recirculation bubble, without acting on

    Fig. 1 aSide view and front view of full-size Ahmed body with 25slant angle.b Schematic view of rear flow topology for slant angle between

    12.5 and 30, from experimental study of Vino et al. (2005)

    1170 Exp Fluids (2012) 52:11691185

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    5/19

    longitudinal structures. According to Thacker et al. (2009),

    expected drag reduction is then about 10%.

    The experimental setup will be described in a first part.

    In the second part, main time-averaged and time-dependent

    results obtained on the base reference Ahmed model will

    be highlighted. The control parametric study using various

    configurations of electrically operated valves in an open-

    loop control scheme will then be presented. The linkbetween the drag reduction and the modification of the

    mean static pressure distribution over the slant and rear

    vertical base of the model is clearly demonstrated.

    2 Experimental setup

    Experiments were carried out in the 5 m 9 3 m test sec-

    tion of the S4 wind tunnel at the Institut AeroTechnique

    (France), using a 1.044-m length Ahmed model mounted

    over a raised floor (see Fig. 2).

    Due to the large cross-section of the wind tunnel com-pared with the model size, no blockage corrections are

    necessary in the present study (blockage ratio B = 0.7%).

    2.1 The Ahmed reference body

    The generic car model used here is the one originally

    described in Ahmed et al. (1984). In the present study, we

    focus on the 25 slanted rear end in order to deal with the

    high-drag configuration characterized by a large separation

    bubble over the slanted surface along with highly energetic

    streamwise vortices created along the slant side edges

    (see Fig. 1b). The dimensions and the overall shape of the

    model are given in Fig. 1a. The main dimensions of the

    model are L = 1.044 m in length, H = 0.288 m in height

    andl = 0.389 m in width. The height of the 25 rear slant

    is h = 0.094 m, and the height of the rear base is

    Hv = 0.194 m.

    2.2 Wind tunnel and reference incoming flow

    Experiments were carried out for flow velocities ranging

    from 20 to 40 m s-1. The Reynolds number,ReL = U0L/t,

    based on the overall lengthLof the model, ranges between1.4 9 106 and 2.8 9 106. The turbulence level in the S4

    wind tunnel is less than 1.2% over this velocity range.

    In order to reduce the influence of the natural boundary

    layer growing on the wind tunnel ground, the Ahmed body

    is fixed over a raised floor, 0.115 m above wind tunnel

    ground. Other dimensions are given in Fig. 3.

    Special attention has been given to the raised floor

    leading edge in order to avoid any boundary layer sepa-

    ration of the incoming flow. It has been designed based on

    a NACA 0018 airfoil (Fig. 3) to avoid the increase in static

    pressure that could generate a massive boundary layer

    separation of the overall wind tunnel boundary layer.This modification significantly decreases the amount

    of perturbations coming to the model and reduces the

    incoming boundary layer thickness.

    2.2.1 Upstream boundary layer

    The incoming boundary layer profile has been measured

    0.7 m downstream from the raised floor leading edge, i.e.,

    DX/L = -0.29 upstream from the Ahmed model. The

    mean velocity profile is reported in Fig. 4for a wind tunnel

    velocity U0 = 20 m s-1. The boundary layer thickness d

    is calculated according to the 99% criterion. The shape

    parameter HBL has been calculated using the classical

    definition of the boundary layer displacement and

    momentum thickness. The boundary layer thickness d is

    about 25 mm, and the shape parameter HBL is close to

    1.25. The incoming boundary layer is then fully turbulent,

    Fig. 2 View of the experimental setup in the 5 m 9 3 m test section

    of the S4 wind tunnel

    Fig. 3 Schematic description of the experimental setup (Ahmed

    model on the raised floor)

    Exp Fluids (2012) 52:11691185 1171

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    6/19

    which is confirmed by the 1/8th power law that fits well the

    experimental measurements.

    2.3 Experimental measurements

    2.3.1 Aerodynamic balance

    Time-averaged forces were measured using a six-compo-

    nent strain gauge balance mounted under the raised floor.

    The balance is located in a dedicated rounded compartment

    to avoid stream-induced perturbations on the force mea-

    surements (Fig. 5).

    Calibration was made out of wind tunnel using standard

    procedure, and calibration checks were also conducted in

    situ. The maximum error in drag measurements associated

    with repeatability and hysteresis was found to be approx-imately 0.5%.

    2.3.2 Wall-pressure measurements

    Steady wall-pressure measurements were carried with 121

    pressure taps located inhomogeneously mainly on the

    slanted surface, as well as on the roof end and vertical rear

    base. Because of the body symmetry, only a half of the

    model was equipped. Symmetry of the flow was previously

    checked with the help of surface oil flow visualizations on

    the entire slant surface. All the taps were plugged in a

    Scanivalve pressure scanner. Precision of this system is

    usually 0.03% of the full scale.

    On the other half of the model, six piezoelectric micro-

    sensors were implanted. These sensors allow both steady

    and unsteady measurements in areas where characteristic

    structures of the Ahmed wake are expected. These sensors

    have a typical sensitivity of 1 Pa. The location of all these

    pressure taps is shown in Fig. 6.

    2.3.3 Surface oil flow visualizations

    Surface oil flow visualizations were conducted using a

    mixture of silicone oil, dodecan, titanium dioxide and oleic

    acid. This mixture was applied with paintbrushes on the

    slanted surface and allows visualization of friction lines

    when the model is exposed to air flow. Friction lines give

    information about mean flow topology (Gillieron2000).

    2.3.4 Near-wake total pressure loss measurements

    Wake measurements were carried out with the help of a

    two-dimensional motorized explorer, allowing measure-ments in a transversal plane with hot-wire or Kiel probe.

    The last one could be used for unsteady total pressure

    measurements thanks to the embedded piezoelectric sensor.

    2.3.5 Description of the pulsed jets control device

    Pulsed jets are obtained using eight electromagnetic binary

    valves feeding a rectangular chamber before flowing

    through a removable perforated plate (Fig. 7).

    Fig. 4 Evolution of the boundary layer on the raised floor upwind

    the Ahmed model for U0 = 20 m s-1 (ReL = 1.4 9 10

    6) at

    DX/L = -0.29

    Fig. 5 Schematic description of the experimental setup (aerodynamic

    forces measurement system)

    Fig. 6 Pressure taps distribution on the rear end of the model, the

    open circles correspond to the 6 piezoelectric sensors

    1172 Exp Fluids (2012) 52:11691185

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    7/19

    Thanks to this setup, the jets geometry and configura-

    tions can be easily changed from a long continuous slot to a

    set of winglet-type jets just by changing the removable

    plate. Electromagnetic valves (Matrix Ltd.,) are built as a

    magnetic circuit closed by a steel spring tongue, which can

    take one of two stable positions. A short, low-energy

    electric impulse applied to the coil can change the spring

    position to the opposite one, thereby clearing or closing theoutput opening. The valve controller is stimulated by a sine

    wave generator furnishing a wavy train of variable fre-

    quency (Fj) in the range of 5300 Hz. The level of the

    pressure impulse, and then the jet velocityUj, can be varied

    by simply changing the pressure supply level (Pj).

    To ensure the spatial homogeneity of the jet speed along

    the actuation slot, small calibrated balls (2 mm diameters)

    are set in the chamber between the valves exits and the

    perforated plate mounted on the wall of the model (Fig. 7).

    The porous layer distributes the air flow from each valves

    exhaust to the entire surface of the perforated plate.

    A typical time history of the jet velocity Uj(t) measuredwith a hot-wire 1 mm above a jet exhaust is shown in

    Fig.8. Due to the valve technology, the jet velocity is

    periodic but not sinusoidal. Indeed, the signal is closer to a

    square wave signal but exhibits a significant overshoot and

    associated rebounds. This overshoot is characteristic of this

    type of valves, which induces brutal pressure release

    immediately after the valve opening.

    In the present paper, the pulsed jets are characterized by

    their mean velocity Uj and main frequency Fj.

    One can notice that this signal exhibits some additional

    fluctuations. It is not clear whether it can have an influence

    on the flow control experiment. This point is complex and

    still under consideration.

    2.3.6 Description of the flow control configurations

    Three different perforated plates have been used in order to

    test the influence of different types of pulsed perturbations

    (Fig. 9): discontinuous slot, continuous slot andwinglets. Each plate can be used on two different locations:

    roof end 100 mm upstream the slant edge (X/L = -0.1),

    and slant top edge 15 mm (X/L = 0.01) downstream the

    slant upper edge (Fig. 9). Blowing sections are detailed in

    blue, with dimensions in millimeters. The choice of these

    configurations is justified in Sect.4.

    3 Characterization of the base flow

    This first step is obviously to study the natural flow aroundthe body. This knowledge will be helpful to compare to

    previous studies and will help in understanding the flow

    control mechanisms.

    3.1 Drag coefficient

    Drag force is expressed by its drag coefficient, with the

    following expression:

    CX FX

    12qSU20

    1

    where FX is the drag force measured by the aerodynamicbalance,q is the air density (corrected with the atmospheric

    pressure and wind tunnel ambient temperature), S is the

    model cross-section (excluding struts) and U0 is the free

    stream velocity.

    This drag coefficient was measured for several Reynolds

    numbers ReL corresponding to U0 from 20 to 40 m s-1.

    Results are shown in Fig. 10.

    A significant Reynolds effect is observed as the drag

    coefficient decreases with increasing Reynolds number

    (from CX = 0.335 at ReL = 1.4 9 106 to CX = 0.312 at

    ReL = 2.7 9 106). However, these results are consistent

    with other studies like Aider et al. (2009) and Roumeas(2006) in the same Reynolds number range.

    3.2 Steady wall-pressure distributions on the rear end

    and associated surface oil flow visualizations

    As the flow is symmetric, steady wall-pressure measure-

    ments were made over the half of the slant surface allowing

    pressure coefficient mappings on the entire slant surface.

    Pressure coefficient Cp is expressed as:

    Fig. 7 Schematic description of the pulsed jets device using perfo-

    rated plates system to change the jets geometry

    Fig. 8 Typical time history of jet velocity Uj(t) (Fj = 200 Hz,

    measured at 1 mm above a jet exhaust)

    Exp Fluids (2012) 52:11691185 1173

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    8/19

    Cp p p0

    12qU20

    2

    where p is the local static pressure and p0 is a reference

    static pressure measured upwind the model, in the undis-

    turbed flow.

    In order to highlight the link between flow structures and

    pressure distribution, pressure coefficient mappings are

    associated with surface oil flow visualizations on Fig. 11.

    From left to right, Reynolds number increases from

    1.4 9 106 (20 m s-

    1) to 2.8 9 106 (40 m s-

    1). XS is localaxis, i.e., the x-axis projection along the slant.

    For eachReynolds number, one can observe a low-pressure

    area on the top of the slant, followed by a gradual pressure

    recovery at the bottom of this surface. Comparing with cor-

    responding flow visualizations, the pressure contours match

    well with the recirculation bubble (circled in red), which is

    responsible for the low-pressure distribution in this area.

    On both sides of the slant, other low-pressure areas are

    visible. They are located under the longitudinal vortices

    (Fig.1b), which also have an important role on the low-pressure repartition and thus on the drag.

    A significant Reynolds effect can be observed on the

    mean pressure distribution (Fig. 11). Indeed, the recircu-

    lation area significantly decreases with the Reynolds

    number. The upper slant low-pressure area then gets

    smaller, and pressure recovery occurs sooner. Even though

    the pressure coefficient becomes smaller on the top of the

    slant, its reduction cannot balance the earlier pressure

    recovery leading to an overall mean pressure value on the

    slant surface, which increases with the Reynolds number.

    This tendency is consistent with the drag reduction

    observed on Fig.10in the same range of Reynolds number.The highest drag configuration that exhibits the largest

    separation bubble (for ReL = 1.4 9 106) has been chosen

    to carry out the flow control experiments.

    3.3 Near-wake total pressure loss measurements

    Base flow topology has been investigated using time-

    averaged total pressure loss measurements in the near

    wake. Results are presented as total pressure loss coeffi-

    cient defined as:

    Cpi 1

    pT p012qU20 3

    where pT is the total pressure measured in the wake. This

    coefficient value is zero in the undisturbed flow (i.e., no

    pressure loss) and gradually increases as total pressure in

    the wake decreases due to pressure losses associated with

    mixing processes (shear layer, recirculation areas,

    vortices ).

    Results are reported on Fig.12 for a cross-section

    located at DX = 0.144 m behind the model (i.e., a relative

    Fig. 9 Perforated plates

    geometry and blowing locations

    Fig. 10 Evolution of the drag coefficient with the Reynolds number

    ReL (without control)

    1174 Exp Fluids (2012) 52:11691185

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    9/19

    distance DX/H = 0.5), for U0 = 20 m s-1 (ReL = 1.4 9

    106). The structure of the wake is classic: the conical

    streamwise vortex signature (with a core located near

    Z1 = 220 mm andY1 = 85 mm (Y1,Z1) being a local axis

    system associated with the plan-wake), the base ring-

    shaped structure and the mixing region associated with the

    flow separation over the slant surface.

    3.4 Velocity profiles

    Various velocity profiles have been measured in several

    locations to complete the characterization of the mean base

    flow.

    3.4.1 Boundary layer over the model roof

    Figure13 shows the boundary layer profile on the model

    roof atX = -0.1 m (X/L& -0.1) upstream from the slant

    edge. The boundary layer thickness isd & 24 mm, and the

    shape parameter HBL & 1.21. The incoming boundary

    layer is then fully turbulent, which is confirmed by the

    1/8th power law that fits well the experimental measure-

    ments.

    3.4.2 Shear layer

    The recirculation bubble is separated from the so-called

    external flow region by a shear layer characterized by a free

    stream velocity (close to the wind tunnel velocity) and a

    Fig. 11 Influence of the Reynolds number on the pressure coefficient distribution on the rear slant and associated surface oil flow visualizations

    (without control)

    Fig. 12 Total pressure loss coefficient distribution in the near wake

    of the model in theDX/H = 0.5 cross-section (without control) in the

    (Y1, Z1) local axis system associated with the plan-wake

    Fig. 13 Boundary layer mean velocity profile on the model roof for

    U0 = 20 m s-1 (ReL = 1.4 9 10

    6) at X/L& -0.1

    Exp Fluids (2012) 52:11691185 1175

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    10/19

    quasi-zero velocity area in the recirculation region. This is

    a region of intense mixing characterized by strong velocity

    gradients and turbulence intensities. Velocity profiles have

    been measured in the vertical symmetry plane (Y = 0) of

    the Ahmed model at two distinct axial positions from the

    slant edge: X/h & 0.1 and X/h & 0.5 (with h, the height

    of the slant face). Results are reported on Fig.14 for

    U0 = 20 m s-1.The mean velocity profiles are well fitted by a hyper-

    bolic-tangent velocity profile. The model used is expressed

    as (Ho and Huerre1984):

    u z U 1 R tanh z z0

    2h

    h i 4

    where U Umax Umin=2 is the average velocity,

    R DU=2Uis the velocity ratio, DU Umax Umin is thetotal shear,z0the mean vertical position of the shear layer,

    i.e., the position of the inflexion point, and h is the

    momentum thickness of the shear layer.

    Values of those parameters for both the positionsX/h & 0.1 and X/h & 0.5 are reported in Table 1, along

    with the associated Reynolds numbers based on the

    momentum thickness and average velocity Reh Uh=m.

    3.5 Unsteady measurements

    The Ahmed model exhibits intense unsteady three-dimen-

    sional wake. According to Thacker (2010), this unsteadi-

    ness is mainly concentrated in the shear flow region over

    the slant surface and in the near-wake region. In the shear

    flow region, the unsteady flow features can be associated

    with both an absolute and convective instability of the

    shear layer (Cherry et al. 1984; Kiya and Sasaki 1985;

    Thacker 2010). The former being associated with a

    flapping of the shear layer and the latter at a natural

    KelvinHelmholtz instability of the shear layer (Aider

    et al. 2007). On the other hand, the unsteady character-

    istic of the near-wake flow region is mainly the conse-

    quence of an unstable organization due to the flow

    separation on both the upper and lower edges of the rearvertical base linked to the ring-shaped structure observed

    in the wake.

    According to Thacker (2010), a significant level of

    velocity fluctuations can also be measured in both the two

    steady streamwise vortical structures, but mainly due to an

    interaction with the unsteadiness of the shear layer, and

    thus only at a significant distance from the core of the

    vortices.

    3.5.1 Unsteady organization of the shear layer

    Unsteady velocity measurements have been performed inthe shear layer for both the positions X/h & 0.1 and

    X/h & 0.5. Power spectral densities associated with

    velocities measured at the inflexion point of both shear

    layer profiles (see on Fig. 14) are reported on Fig. 15.

    Results are shown in a non-dimensional form introducing

    the Strouhal number Sth = fh/U0 (the reduced frequency

    basedon the slant height),for two wind tunnelvelocitiesin

    order to highlight specific unsteady organization that

    could be characterized by a constant value of Strouhal

    number.

    At the position X/h & 0.1, the non-dimensional spec-

    trums exhibit a significant low-frequency organization

    Fig. 14 Shear layer velocity

    profiles, with:a mean velocity

    profiles andb root mean square

    velocity profiles, at two axial

    positions from the slant edge:

    X/h & 0.1 andX/h & 0.5, for

    U0 = 20 m s-1

    (ReL = 1.4 9 106)

    1176 Exp Fluids (2012) 52:11691185

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    11/19

    characterized by a Strouhal number Sth & 0.1 for

    U0 = 20 m s-1 and Sth & 0.14 for U0 = 30 m s

    -1, with

    frequency values of, respectively, f& 20 Hz and

    f& 45 Hz. According to Kiya and Sasaki (1985), a con-

    stant Strouhal number can be associated with this low-

    frequency organization due to the flapping of the shear

    layer, with an appropriate definition of the reduced fre-

    quency based on the length of the recirculation bubble Lr.

    Results are reported in Table2, where Lr has been esti-

    mated from the surface oil flow visualizations (see on

    Fig.11).

    At the position X/h & 0.5, the non-dimensional spec-

    trums exhibit the same low-frequency organization along

    with a significant increase in the energy fluctuations in a

    higher frequency range between Sth & 0.5 (f& 100 Hz

    for U0 = 20 m s-1 and f& 150 Hz for U0 = 30 m s

    -1)

    and Sth & 2 (f& 450 Hz for U0 = 20 m s-1 and

    f& 650 Hz for U0 = 30 m s-1). Taking the shear layer

    relevant parameters, i.e., the momentum thickness h and

    the average velocity U measured at X/h & 0.5 (see

    Table1), those high-frequency fluctuations occur between

    Sth & 0.033 andSth & 0.13. In the light of the work of Ho

    and Huerre (1984), those fluctuations can then be associ-

    ated with the roll-up of the shear layer due to the Kelvin

    Helmholtz instability mechanism.

    3.5.2 Unsteady organization of the near wake

    Unsteady velocity measurements have been also carried

    out on several locations in the near wake, exhibiting a

    strong unsteady organization characterized by a constant

    Strouhal numberStHv & 0.31 (based on the rear-end ver-

    tical height Hv = 0.194 m).

    Results are reported on Fig. 16 for a point near the

    bottom of the rear end where the organization is particu-

    larly strong (Z/H = -1). Indeed, the non-dimensional

    spectrums (based on the power spectral densities of mea-

    sured velocities) exhibit strong and narrow peaks for a

    reduced frequency StHv & 0.31, indicating a very orga-

    nized phenomenon. Taking another definition of the

    reduced frequency based on the square root of the model

    cross-section A =HS, one find StA = fA/U0 & 0.53,which is in accordance with the results of Vino et al. ( 2005)

    and Thacker (2010).

    Table 1 Parameters of the

    hyperbolic-tangent velocity

    profiles used atX/h & 0.1

    and X/h & 0.5

    X/h Umax (m s-1) Umin (m s

    -1) U R Z0 (mm) h (mm) Reh

    0.1 21 0 10.5 1 -2.75 1.2 700

    0.5 20.5 3.5 12 0.71 -12 3.75 3,000

    Fig. 15 Non-dimensionalpower spectral densities of the

    velocity at the inflection point of

    the shear layer at a X/h & 0.1

    andb X/h & 0.5 for

    U0 = 20 m s-1 and 30 m s-1

    (ReL = 1.4 9 106 and

    ReL = 2.1 9 106)

    Table 2 Parameters of the low-frequency organization of the shear

    layer forX/h & 0.1

    U(m s-1) f(Hz) Sth Lr (m) StLr

    20 &20 &0.1 &0.17 &0.1730 &45 &0.14 &0.13 &0.195

    Exp Fluids (2012) 52:11691185 1177

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    12/19

    4 Flow control experiments

    Among the control parameters, one can identify physical

    parameters associated with the jet (mean and maximum jet

    velocities, pulsation frequency, duty cycle, signal form,

    etc.) and geometric parameters (shape of the cross-section

    of the nozzle, number and spatial organization of the jets,

    location of the jets over the model, jets angles, etc.) In the

    present work, control experiments were realized with four

    different geometric configurations:

    Discontinuous slot, corresponding to rectangular jets

    aligned along the spanwise direction, over the slant

    upper edge and roof end.

    Continuous open slot close to the slant upper edge.

    Winglets jets over the roof end.

    Each of them corresponds to different flow control

    strategies.

    The discontinuous slot and winglets configurations over

    the roof end correspond to jet vortex generators. The

    objective is to test, with two geometrical configurations,

    the impact of longitudinal vorticity injection upstream the

    separation point. The idea is to use pairs of streamwise

    counter-rotating vortices induced by the jets in cross-flow

    (Cortelezzi and Karagozian 2001) to modify the property

    of the boundary layer and postpone the separation of the

    boundary layer (Duriez et al. 2006,2008a,b). In the case of

    pulsed jets, one can expect both a modification of the mean

    flow and, as a consequence, of the shear layer, together

    with an effect of the pulsation frequency injected in the

    shear layer.

    The continuous slot at slant top edge was intended to

    quantify the effect of transversal vorticity injection near the

    separation point. This configuration is inspired by the work

    of Leclerc (2008) who showed that it was possible to

    decrease the drag with synthetic jets. In this case, the

    injection is homogenous along the spanwise direction so

    that the shear layer is perturbed by a time-periodic span-

    wise vorticity sheet. In this case, no streamwise vorticity is

    injected.

    The discontinuous slot at the same location was used to

    experiment the effect of the reduction of injected

    momentum quantity and spanwise modulation. In this case,

    the shear layer is no longer perturbed by a spanwise vortex

    but rather by a set of streamwise vortices spaced along the

    spanwise direction, even if their location downstream the

    separation make the comparison with jets in cross-flow

    more difficult.

    Dimensionless quantities are used for jets speeds and

    jets frequencies, with the classical definition for momen-

    tum coefficient Cl and dimensionless frequency Stj (Sj is

    the perforated plates blowing surface for the considered

    control configurations):

    Cl qSjU

    2j

    1=2qSU20

    5

    Stj Fjh

    U06

    For each geometric configuration, two physical para-

    meters, jets speed and jets frequency, were varied. For

    every parameter, drag coefficients and pressure coefficients

    were measured and plotted as iso-contours in the (Cl,Stj)

    space. White areas in the contour plots correspond to

    parameters that have not been measured mainly because ofelectric power limitations. In the following, all data

    were obtained at ReL = 1.4 9 106 (corresponding to

    U0 = 20 m s-1).

    4.1 Influence of the forcing parameters on drag

    Drag coefficient without and with control are, respectively,

    noted CX0 and CXC. Figures17, 18, 19 and 20 show the

    drag coefficient variations DCX = (CX0 - CXC)/CX0 as a

    function of the momentum coefficient Cl and dimension-

    less frequency Stj for the four geometric configurations. In

    the following, the space parameter is mapped with incre-

    ments dFj = 20 Hz or dFj = 40 Hz (i.e., dStj & 0.1 or

    dStj & 0.2) depending on the tested configuration. In the

    same way, Cl variations were obtained by changing Pjwith dPj = 0.5 bar and dPj = 1 bar, corresponding to

    dCl = 0.3 9 10-3 or dCl = 0.6 9 10-3. DCX[ 0 cor-

    responds to drag reduction in percentage.

    One can observe that 8% drag reduction is reached for

    various configurations and that the corresponding physical

    parameters depend strongly on the geometric ones.

    Fig. 16 Power spectral

    densities of the velocity in the

    near wake forU0 = 20 m s-1

    and 30 m s-1

    (ReL = 1.4 9 106 and

    ReL = 2.1 9 106)

    1178 Exp Fluids (2012) 52:11691185

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    13/19

    The discontinuous slot at roof end (Fig. 17) leads to drag

    reductions for almost all the frequency range tested with

    two optimal Strouhal number at Stj = 0.62 andStj = 1.03.

    This drag reduction seems to be optimal for a relatively

    narrow range of momentum coefficients 3 9 10-3\Cl\

    3.5 9 10-3. For this configuration, both the amount of

    injected momentum quantity and pulsation frequency play

    an important role, even if the influence of the pulsation

    frequency is weaker.

    On the opposite, the continuous slot configuration at

    slant top edge (Fig.18) leads to a significant drag reduction

    for a particular frequency range 0.3\ Stj\0.6 with

    local optimal areas for Stj & 0.28 and Cl & 2.2 9 10-3,

    Stj & 0.35 and 0.4 9 10-3\Cl\0.9 9 10-3 and

    Stj & 0.56 andCl & 0.3 9 10-3. The influence of the jet

    velocity (and then of the injected momentum) seems here

    to be linked to the jet pulsation frequency that plays a

    major role in the drag reduction mechanism.

    Two important points should be noticed. First, the

    evolution of the drag reduction in the (Cl, Stj) space

    parameters is completely different for the two configura-

    tions. It confirms that the actuations (and control strategy)are completely different. Second, the maximum drag

    reduction is the same (about 8%) but for much smaller Cl

    (0.4 9 10-3 instead of 3 9 10-3) with the continuous slot

    at slant edge.

    It is not possible to link the optimal frequency for the

    drag reduction to natural frequencies measured in the shear

    layer region. Nevertheless, one can notice that some of

    them are close to the KelvinHelmholtz frequency

    Sth & 0.5 measured at X/h & 0.5 in the shear layer.

    For the two last configurations (winglets at roof end on

    Fig.19and discontinuous slot at slant top edge on Fig. 20),

    the drag reduction is maximal for given points in the spaceparameter: Stj & 1.1 and Cl C 1.2 9 10

    -3 for the wing-

    lets at roof end configuration and 1 9 10-3\Cl

    \ 2.2 9 10-3 and Stj & 0.55 or Stj C 0.9 for the discon-

    tinuous slot at slant top edge configuration. In these cases,

    both jet velocity and pulsation frequency seem to be

    important to optimize the drag reduction. In the case of the

    discontinuous slot at slant top edge, the optimal frequency

    is also close to the natural frequency measured in the shear

    layer.

    Fig. 17 Influence of the control parameters (Stj and Cl) on the drag

    reduction (in %) for the discontinuous slot at roof end configuration at

    U0 = 20 m s-1 (ReL = 1.4 9 10

    6)

    Fig. 18 Influence of the control parameters (Stj and Cl) on the drag

    reduction (in %) for the continuous slot at slant top edge end

    configuration atU0 = 20 m s-1 (ReL = 1.4 9 10

    6)

    Fig. 19 Influence of the control parameters (Stj and Cl) on the drag

    reduction (in %) for the winglets at roof end configuration atU0 = 20 m s

    -1 (ReL = 1.4 9 106)

    Fig. 20 Influence of the control parameters (Stj and Cl) on the drag

    reduction (in %) for the discontinuous slot at slant top edge

    configuration atU0 = 20 m s-1 (ReL = 1.4 9 10

    6)

    Exp Fluids (2012) 52:11691185 1179

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    14/19

    By comparing the Figs. 17and 19, and Figs. 18 and 20,

    one can observe that changing the jet exhaust geometry for

    a fixed forcing location dramatically changes the control

    behavior. Changing discontinuous slot for winglets on the

    roof position suppresses the local optimum of drag reduc-

    tion identified at Stj = 0.62, and exchanging continuous

    slot by discontinuous slot also suppresses several optimal

    frequencies like Stj = 0.35.It is also interesting to notice that the most efficient

    frequency is about two times the natural frequency of the

    shear layer when the perturbation is located upstream of the

    separation, while it is close to the natural frequency when

    the perturbation is close and downstream of the separation.

    Table3 summarizes the better drag reductions and

    corresponding parameters for each configuration. The

    continuous slot at slant edge is clearly the most efficient

    configuration with a much smaller optimal Cl. From the

    industrial point of view, it is also important to notice that

    most of the perturbations lead to significant drag reduction

    so that it is possible to choose the right flow controlstrategies depending on the location where it can be inte-

    grated in the vehicle.

    4.2 Influence of the forcing parameters on local

    pressure

    In order to explain the drag reductions, mean local static

    pressure has been monitored at various locations using

    pressure sensors. Only a few sensors have been used during

    control tests, so those measurements can only give general

    trends. Static pressure results are presented in the same waythan previous drag reduction measurements: for each pair

    of parameters, the corresponding local static pressure var-

    iation DCp = (Cp0 - CpC)/Cp0 is reported. Pressure coef-

    ficients without and with control are, respectively, noted

    Cp0 and CpC.

    4.2.1 Wall pressure at slant upper edge:

    Figure21 presents slant upper edge local pressure varia-

    tions for the same parameters and configurations as the one

    presented on Figs.17,18,19 and 20 for the drag. One can

    observe strong similarity with the drag variation for thediscontinuous slot configurations at slant top edge (by

    comparing Fig. 21a with Fig. 17) and roof end (by com-

    paring Fig.21d with Fig. 20): graphics are nearly identical.

    This means that, for both those configurations, the drag

    reduction is strongly connected with pressure recovery in

    this particular area.

    For the winglets configuration, pressure variation results

    are also quite similar to the drag variation results but with

    lower pressure variations. For the continuous slot config-

    uration at slant top edge, similarity can only be detected

    for two local areas Stj & 0.6/Cl & 0.3 9 10-3 and

    Stj & 1/Cl & 0.5 9 10-3 but with small pressure variations.

    Table 3 Better drag coefficient reductions for each configuration and

    associated parameters

    Pulsed jets configuration DCX (%) Stj Cl (10-3)

    Discontinuous slotroof 7.8 1.03 3.1

    Continuous slotslant edge 7.5 0.35 0.3

    Wingletsroof 6.9 1.13 1.8

    Discontinuous slotslant edge 6.3 1.22 1.7

    Fig. 21 Influence of the control

    parameters (Stj and Cl) on the

    pressure coefficient evolution

    (in %) on the top of the rear

    slant, for the four flow control

    configurations

    1180 Exp Fluids (2012) 52:11691185

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    15/19

    4.2.2 Wall pressure in the middle of the rear slant

    Figure22describes local pressure variations at the centerof the slant. Here again, comparison with the drag reduc-

    tion results highlights very similar trends and relation

    between drag and wall-pressure variations. The two dis-

    continuous slot configurations (Fig. 22a, d) and the wing-

    lets configuration (Fig.22b) show areas with pressure

    benefits larger than those observed on drag (3040%). The

    pulsed blowing through continuous slot configuration

    (Fig.22c) also shows areas of pressure variations of almost

    15% for frequency close to Stj & 0.6.

    These measurements confirm that the discontinuous slot

    at roof end reduces the drag through pressure recovery over

    the rear slant. As expected, drag reductions with the otherconfigurations also induce an increase in the pressure dis-

    tribution over the rear slant.

    4.2.3 Wall pressure in the middle of the rear end

    Rear blunt local pressure variations are plotted in Fig. 23.

    Here again, comparison with drag reduction mappings

    in Figs. 17, 18, 19 and 20 enables to locate where a par-

    ticular configuration produces benefits. Except for the

    Fig. 22 Influence of the control

    parameters (Stj and Cl) on the

    pressure coefficient evolution

    (in %) on the middle of the rear

    slant, for the four flow control

    configurations

    Fig. 23 Influence of the control

    parameters (Stj and Cl) on the

    pressure coefficient evolution

    (in %) on the middle of the rear

    end, for the four flow control

    configurations

    Exp Fluids (2012) 52:11691185 1181

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    16/19

    discontinuous slot at roof end configuration (Fig. 23a), all

    other three configurations (Fig.23b, c, d) show similar

    trends than the drag reduction results. For those specific

    configurations, the drag reduction can then be linked with

    the pressure recovery over the rear-end surface of the body.

    4.2.4 Conclusion on local pressure variation results

    In light of previous results, it is now possible to clarify the

    influence of each particular control strategies on the nearwake of the Ahmed body:

    The pulsed blowing through the discontinuous slot at

    roof end has a strong influence on the wall pressure in

    the top and middle of the slant. This suggests an action

    mainly on the recirculation bubble.

    The continuous slot at slant top edge seems mainly to

    modify the wall pressure on the middle of the slant and

    on the rear blunt part of the model, suggesting a

    modification of the ring-shaped vortical structure.

    The pulsed blowing through winglet-type jets acts on

    the wall pressure over the slant with a strongest effectnot only in the middle of the slant but also in the rear

    blunt part of the model. One can hypothesize a

    modification of the shear layer starting from the top

    of the slant and, as a consequence, a stronger interac-

    tion between the shear layer and the torodal recircu-

    lation on the rear part of the model.

    As for the winglets configuration, the discontinuous slot

    at roof end modifies the mean pressure value at each

    location (top and middle of the slant and rear vertical

    surface), suggesting a complex interaction between the

    main structures of the wake.

    In order to highlight previous results, the near-wake

    modification has been investigated for the best control

    strategy, i.e., the pulsed blowing through the discontinuous

    slot at roof end (Fig. 17), and compared with the base flow

    without control.

    4.3 Near-wake modification by pulsed blowing

    Time-averaged total pressure loss coefficient mappings

    in a vertical cross-section located at a relative distance

    DX/H = 0.5 behind the model for ReL = 1.4 9 106 are

    presented on Fig. 24, for both the base flow and the con-

    trolled flow. The controlled flow result is the one associated

    with the discontinuous slot at roof end, with the better set

    of parameters defined in Table 3.

    One can notice several differences between the natural

    and the controlled flows:

    The total pressure loss coefficient area associated with

    the slant recirculation bubble is clearly reduced, whichconfirms the effectiveness of the pulsed blowing on the

    roof end boundary layer separation and then on the

    recirculation bubble. It also confirms that the recircu-

    lation bubble is reduced when the pulsed blowing

    reduces the drag.

    As expected (Aider et al.2009; Fourrieet al.2011), the

    control system acts also on the longitudinal vortices: a

    pressure loss drop happens at the edge of the structure,

    while losses in the core seem to weakly increase. One

    Fig. 24 Half planes of total pressure loss coefficient distributions in the near wake of the modela without control,b with control andc the Cpidifference between the two cases

    1182 Exp Fluids (2012) 52:11691185

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    17/19

    can hypothesize that this is an effect of the complex

    interaction existing between separation bubble and

    longitudinal structures: the reduction of the former

    allows the later to develop with a smaller size but a

    stronger intensity.

    Pressure losses become also a bit more important in the

    blunt recirculation. Here again, it is probably a

    consequence of the interaction with the slant separa-

    tion. Various authors (Roumeas et al.2008; Pujals et al.

    2010) report that the cores of the blunt recirculation

    move downstream when the slant recirculation issuppressed: one can made the assumption that blunt

    recirculation cores come closer to the measurement

    plane when control is activated and the recirculation

    bubble suppressed, with the result of more apparent

    pressure losses.

    The standard deviations associated with the unsteady

    total pressure coefficients are plotted on Fig. 25. Without

    control, separation bubble exhibits strong pressure varia-

    tions (Fig. 25a). In the controlled case, fluctuations only

    remain on the longitudinal structures and, even if reduced,

    on the underbody flow (Fig. 25b).

    5 Conclusions

    Time-averaged and time-dependent base flow around a

    standard Ahmed body with 25 slant angle has been

    characterized in wind tunnel. Mean flow and drag results

    are in accordance with previous studies. A significant

    Reynolds effect has been observed in both the drag

    coefficient and mean pressure distribution on the rear end,

    due to a reduction of the recirculation bubble.

    Unsteady measurements in the rear-end flow reveal

    three mechanisms that can be characterized by a constant

    value of reduced frequency. The more organized, linked

    with the ring-shaped structure observed in the wake, is

    characterized by a Strouhal number (based on the rear

    vertical height), StHv & 0.31.

    In the shear flow region that separates the recirculation

    bubble from the external flow region, two unsteady orga-

    nizations have been highlighted. One is characterized by aStrouhal number (based on the slant height) Sth & 1.2 and

    is associated with the natural KelvinHelmholtz instability

    of the shear layer. The other is due to the flapping of the

    shear layer and is characterized by a Strouhal number

    (based on the length of the recirculation bubble)

    StLr & 0.17 at a Reynolds number ReL = 1.4 9 106. At

    this specific Reynolds number, the Ahmed model exhibits a

    high-drag coefficient characterized by a large separation

    bubble along with energetic streamwise vortices. This

    Reynolds number has then been chosen to carry out flow

    control experiments focused on slant recirculation, without

    any attempt to control longitudinal structures.The influence of rear-end periodic forcing on the drag

    coefficient has then been investigated using electrically

    operated magnetic valves in an open-loop control scheme.

    Four distinct configurations of flow control have been

    tested: pulsed jets in a discontinuous slot or in winglets on

    the roof end and in a discontinuous or continuous slot at the

    top of the rear slant. For each configuration, the influence

    of the forcing parameters (non-dimensional frequency,

    injected momentum quantity) on the drag reduction has

    Fig. 25 Half planes of total

    pressure loss coefficient

    fluctuations in the near wake of

    the modela without control and

    b with control

    Exp Fluids (2012) 52:11691185 1183

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    18/19

    been examined, along with their impact on the static pressure

    on both the rear slant and vertical base of the model.

    Maximum reductions between 6 and 8% have been

    measured depending on the geometric and jet exhaust

    configurations that show different sensitivity to the forcing

    parameters.

    Indeed, for the jets pulsing through discontinuous slots at

    roof end, the amount of injectedmomentum quantity seems tobe the key parameter, and there is a weaker influence of the

    pulsation frequency. On the opposite, the vortical sheet puls-

    ing through continuous slot at slant top edge leads to a sig-

    nificant drag reduction only for a particular frequency range.

    For the two last configurations (jets pulsing through winglets

    at roof end or through discontinuous slots at slant top edge),

    bothjet velocity and pulsation frequency seemto be important

    to optimize the drag reduction.

    However, the influence of other parameters has to be

    investigated. The control of the amplitude of the jet

    velocity signal is probably an important feature, while the

    study of the duty cycle may be a promising way in order toreduce the needed momentum quantity. In the present

    study, an important overshoot has been observed, and it

    would be interesting to highlight its effect on the flow

    control.

    One can also notice that when the perturbations are

    close to the separation the blowing frequencies that pro-

    duce the best results are close to the KelvinHelmholtz

    instability frequency of the shear layer or between the

    flapping frequency and the KelvinHelmholtz instability

    frequency of the shear layer. This is in accordance with the

    work of Sigurdson (1995) on the effect of a periodic

    velocity perturbation on the separation bubble downstream

    of the sharp-edged blunt face. On the contrary, the most

    efficient frequencies when the flow is perturbed upstream

    the separation are about two times the natural shedding

    frequency. This is of course different from the work of

    Sigurdson where there is no incoming boundary layer.

    Meanwhile, further investigations need to be done to

    highlight in our case the impact of the velocity perturbation

    on the entrainment of flow and/or growth rate of the shear

    layer and the impact on the reattachment length.

    Acknowledgments This work was carried out in the framework of

    the CARAVAJE project supported by the Agence pour le Devel-oppement Et la Matrise de lEnergie (ADEME). We thank the

    Renault SA and PSA Peugeot-Citroen Aerodynamics Research teams

    and the Plastic Omnium research team for fruitful discussions.

    Technical support by the S4 Wind Tunnel team is also gratefully

    acknowledged.

    References

    Ahmed SR, Ramm G, Faltin G (1984) Some salient features of the

    time-averaged ground vehicle wake. SAE 840300

    Aider JL, Danet A, Lesieur M (2007) Large-eddy simulation applied

    to study the influence of upstream conditions on the time-

    dependant and averaged characteristics of a backward-f acing

    step flow. J Turbul 8:N51

    Aider JL, Beaudoin JF, Wesfreid JE (2009) Drag and lift reduction of

    a 3D bluff-body using active vortex generators. Exp Fluids

    48:771789

    Barnard RH (1996) Road vehicle aerodynamic design: an introduc-

    tion. Longman, Essex

    Beaudoin JF, Aider JL (2008) Drag and lift reduction of 3D bluff-

    body using flaps. Exp Fluids 44:491501

    Beaudoin JF, Aider JL, Cadot O, Gosse K, Paranthoen P, Hamelin B,

    Tissier M, Wesfreid JE (2004) Characterization of longitudinal

    vortices on a 3D bluff-body using cavitation. Exp Fluids

    37:763768

    Beaudoin JF, Cadot O, Wesfreid JE, Aider JL (2008) Feedback

    control using extremum seeking method for drag reduction of a

    3D bluff body. In: IUTAM symposium on flow control and

    MEMS, London

    Brunn A, Nitsche W, Henning L, King R (2008) Application of

    Slope-seeking to a Generic Car Model for Active Drag Control.

    In: 26th AIAA applied aerodynamics conference, Honolulu

    Cherry NJ, Hillier R, Latour MEMP (1984) Unsteady measurements

    in a separated and reattaching flow. J Fluid Mech 144:1346

    Cortelezzi L, Karagozian AR (2001) On the formation of the

    counter-rotating vortex pair in transverse jets. J Fluid Mech

    446:347373

    Duriez T, Aider JL, Wesfreid JE (2006) Base flow modification by

    streamwise vortices. Application to the control of separated

    flows. In: ASME Joint U.S.European fluids engineering

    summer meeting, Miami

    Duriez T, Aider JL, Wesfreid JE (2008a) Control of a separated flow

    over a smoothly contoured ramp using vortex generators. In:

    IUTAM symposium on flow control and MEMS, London

    Duriez T, Aider JL, Wesfreid JE (2008b) Non-linear modulation of a

    boundary layer induced by vortex generators. AIAA 2008-4076

    Fares E (2006) Unsteady flow simulation of the Ahmed reference

    body using a lattice Boltzmann approach. Comput Fluids

    35:940950

    Fourrie G, Keirsbulck L, Labraga L, Gillieron P (2011) Bluff-body

    drag reduction using a deflector. Exp Fluids 50:385395

    Gillieron P (2000) La technique des visualisations parietales. Lesson,

    Conservatoire National des Arts et Metiers

    Gillieron P (2010) Influence of the slant angle of 3D bluff-bodies

    on longitudinal vortex formation. J Fluids Eng 132(0511041):

    051104051109

    Ho CM, Huerre P (1984) Perturbed free shear layers. Ann Rev Fluid

    Mech 16:365424

    Hucho WH (1998) Aerodynamics of road vehicles. Cambridge

    University Press, Cambridge

    Kiya M, Sasaki K (1985) Structure of large scale vortices and

    unsteady reverse flow in the reattaching zone of a turbulent

    separation bubble. J Fluid Mech 154:463491

    KrajnovicS, Davidson L (2005a) Flow around a simplified car, part 1:large eddy simulation. J Fluids Eng 127:907918

    KrajnovicS, Davidson L (2005b) Flow around a simplified car, part 2:

    understanding the flow. J Fluids Eng 127:919928

    Krajnovic S, Osth J, Basara B (2009) LES of active flow control

    around an Ahmed body with active flow control. In: Conference

    on modelling fluid flow (CMFF09), Budapest

    Krentel D, Mumiovic R, Brunn A, Wolfgang N, King R (2010)

    Application of active flow control on generic 3D car models. In:

    King R (ed) Active flow control II 2010. Springer, Berlin

    Leclerc C (2008) Reduction de la trainee dun vehicule automobile

    simplifie a laide du controle actif par jet synthetique. PhD

    thesis, Institut National Polytechique de Toulouse

    1184 Exp Fluids (2012) 52:11691185

    1 3

  • 8/9/2019 25 Degree Ahmed Body aerodynamic study

    19/19

    Lehugeur B (2009) Controle des structures tourbillonnaires longitu-

    dinales dans le sillage dune geometrie simplifiee de vehicule

    automobile: approche experimentale. Mech Ind 9:533541

    Pastoor M, Henning L, Noack BR, King R, Tadmor G (2008)

    Feedback shear layer control for bluff body drag reduction.

    J Fluid Mech 608:161196

    Pujals G, Depardon S, Cossu C (2010) Drag reduction of a 3D bluff-

    body using coherent streamwise streaks. Exp Fluids 49:1085

    1094

    Roumeas M (2006) Contribution a lanalyse et au controle des

    sillages de corps epais par aspiration ou soufflage continu. PhD

    thesis, Institut National Polytechnique de Toulouse

    Roumeas M, Gillieron P, Kourta A (2008) Drag reduction by flow

    control on a car after body. Int J Numer Meth Fluids 60:1222

    1240

    Sigurdson LW (1995) The structure and control of a turbulent

    reattaching flow. J Fluid Mech 298:139165

    Thacker A (2010) Contribution experimentale alanalyse stationnaire

    et instationnaire de lecoulement alarriere dun corps de faible

    allongement. PhD thesis, UniversitedOrleans

    Thacker A, Leroy A, Aubrun S, Loyer S, Devinant P (2009)

    Caracteristiques du sillage du corps de Ahmed: effet de la

    suppression du decollement de lunette arriere. In: GDR 2502

    Controle des decollements, Orleans

    Tihon J, Penkavova V, Pantzali M (2010) The effect of inlet

    pulsations on the backward-facing step flow. E J Mech B Fluids

    29:224235

    Vino G, Watkins S, Mousley P, Watmuff J, Prasad S (2005) Flow

    structures in the near wake of the Ahmed model. J Fluids Struct

    20:673695

    Exp Fluids (2012) 52:11691185 1185

    1 3