Flow Structure Around A Low-Drag Ahmed Body
Flow Structure Around A Low-Drag Ahmed Body
1136
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
The wake of an Ahmed body may be divided into high- and low-drag regimes where the
rear slant angle (ϕ) is in the ranges of 12.5°–30° and larger than 30°, respectively. This
work aims to gain a relatively thorough understanding of unsteady predominant coherent
structures around an Ahmed body of ϕ = 35° in the low-drag regime. Extensive hot-wire,
wall pressure, flow visualization and particle image velocimetry measurements have been
conducted at Reynolds number Re ∈ [0.3, 2.7] × 105 , based on the square root of the
model’s frontal area. A total of five distinct Strouhal numbers have been identified in the
wake. One of them, Stw ≈ 0.30, is captured behind the vertical base, which is associated
with the structures that emanate from the upper recirculation bubble and pinch off from
the lower bubble, respectively. It is found that Stw scales with a characteristic length αS,
which reflects physically the bubble size, and the Strouhal number Stw+ based on αS is a
constant 0.20, irrespective of the value of ϕ. A corner vortex rolling upstream is observed
near the lower end of the slanted surface, whose formation mechanism and dynamical role
are discussed. The Reynolds-number effect on the flow is also documented. Based on the
present and previously reported data, a conceptual flow structure model is proposed for a
low-drag Ahmed body, including both steady and unsteady coherent structures around the
body.
Key words: separated flows, vortex interactions, wakes
1. Introduction
The environmental issue of air pollution has forced the governments of many countries to
https://doi.org/10.1017/jfm.2020.1136
implement more and more stringent regulations on vehicle emissions. This fact, along with
high fuel costs, highlights the necessity and urgency of searching for new technologies to
reduce aerodynamic drag and hence the fuel consumption of vehicles. The automobile
industry has set an ambitious target to cut down the aerodynamic drag of vehicles by at
characteristics of flow around an Ahmed model with ϕ = 35°. Lienhart et al. (2002) and
Lienhart & Becker (2003) investigated experimentally the time-averaged wake structure,
as well as the surface pressure distribution on the rear window and vertical base
(Re = 8.93 × 105 ). A number of numerical investigations were launched for this flow at
the same Re, focusing on comparison in the time-averaged flow fields between numerical
and experimental results (Fares 2006; Guilmineau 2008). Guilmineau et al. (2017) studied
913 A21-2
Flow structure around a low-drag Ahmed body
numerically the flow around an Ahmed body with ϕ = 35° at Re = 8.93 × 105 using the
improved delayed detached-eddy simulation method. A small secondary recirculation at
the end of the rear window was observed, which Lienhart & Becker (2003) failed to detect
experimentally. Wang et al. (2013) investigated the wake of an Ahmed body with ϕ = 35°
based on extensive particle image velocimetry (PIV) measurements at Re = 0.62 × 105 .
They found that the shear layer over the roof separated at the upper edge of the slant and
was embedded with alternately signed longitudinal vortices. The separated flow merged
into the upper recirculation behind the base. The gap flow separated from the lower edge
of the base was partially drawn into the lower bubble, and partially rolled up to form a
spanwise roll, which was wrapped by longitudinal structures. A conceptual flow structure
https://doi.org/10.1017/jfm.2020.1136
model was proposed for the low-drag Ahmed body wake. However, this model fails to
provide a full picture of the flow. Firstly, it is largely based on the mean flow data,
containing little information on unsteady flow structure, let alone predominant frequencies
in the wake. Secondly, some important details of the coherent structures are missing,
including the development of longitudinal trailing vortices behind the vertical base.
Unsteady structures around a low-drag Ahmed body and the corresponding St have
received some attention. Kohri et al. (2014) measured the predominant frequencies in the
913 A21-3
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
wakes of Ahmed bodies with ϕ = 32.5° and 35° at Re = 0.73 × 105 and captured three
different St, i.e. St ≈ 0.10 in the symmetry plane behind the vertical base, 0.14 in the flow
separated from the sidewall, and 0.27 downstream of the lower edge of the base. Tunay et
al. (2014) detected St = 0.26 in the downwash flow of the wake and 0.09 behind the lower
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
edge of the base (ϕ = 35°, Re = 0.17 × 105 ). Although the two frequencies are nearly the
same as those (St ≈ 0.27 and 0.1) reported by Kohri et al. (2014), the two studies held
different views on their mechanisms. Kohri et al. (2014) connected St ≈ 0.1 and 0.27 to
the downwash flow from the slanted surface and the upwash flow from the lower edge of
the base, respectively. But Tunay et al. (2014) ascribed them to vortex shedding from the
lower base edge and the upper edge of the rear window, respectively. Apparently, there is
a need to clarify the St values as well as their origins. Furthermore, many other important
issues remain to be resolved. For example, how do the unsteady structures interact with
each other? What is the Re effect on the structures?
This work sets out to address the issues raised above based on extensive hot-wire,
pressure, PIV and flow visualization measurements. Experimental details are provided
in § 2. The predominant flow structures captured are presented in § 3, including their St
values. The spatial extents of these structures or St values are discussed in § 4, followed
by the examination of the Reynolds-number effect in § 5. A conceptual model of the flow
structure around a low-drag Ahmed body is proposed in § 6. The work is concluded in § 7.
2. Experimental details
2.1. Experimental set-up
Experiments were carried out in two different closed-circuit wind tunnels, one at Harbin
Institute of Technology, Shenzhen (HITsz), and the other at the Hong Kong Polytechnic
University (HKPU), their working sections being 1.0 m × 0.8 m by 5.6 m long, and
0.6 m × 0.6 m by 2.4 m long, respectively. The flow non-uniformity was 0.1 % and the
longitudinal turbulence intensity was no more than 0.4 % in the test section for both
tunnels. The experimental set-ups were essentially the same, as shown schematically in
figure 1(a). A flat plate with a clipper-built leading edge was installed horizontally above
the floor of the test section, as a raised floor to control the boundary layer thickness. The
dimensions of the plates were 2.6 m × 0.78 m × 0.015 m and 2.2 m × 0.59 m × 0.015 m,
with clearances between the plate and the tunnel floor of 0.275 m and 0.088 m, in the
HITsz and HKPU wind tunnels, respectively.
The vehicle model was a one-third-scaled Ahmed body (figure 1b) with ϕ = 35°,
corresponding to a low-drag body. The full-scale model was described by Ahmed et al.
(1984). The model had an overall length of L = 0.348 m, a width of W = 0.130 m and
a height of H = 0.096 m, supported by four cylindrical struts of 16.67 mm in length
and 10 mm in diameter. The front end of the model was 0.3 m downstream of the
floor leading edge (figure 1a), where the boundary layer thickness was about 4 mm at
U∞ = 8.33 m s−1 or Re = 0.62 × 105 . The blockage ratios of the frontal surface of the
model to the rectangular test section above the raised floor were approximately 2.2 % and
https://doi.org/10.1017/jfm.2020.1136
4.1 % in the HITsz and HKPU wind tunnels, respectively. Figure 1(b) shows the definition
of the right-handed Cartesian coordinate system (x, y, z), with the origin O at the midpoint
of the lower edge of the model vertical base. In this paper, the instantaneous velocity
components in the x, y and z directions are defined as U, V and W, respectively, which
can be decomposed into a time-averaged component and a fluctuating component, i.e.
U = Ū + u, V = V̄ + v and W = W̄ + w, where overbar denotes time averaging and u, v
and w are fluctuating velocities.
913 A21-4
Flow structure around a low-drag Ahmed body
(a)
Traversing mechanism
U∞
Hot-wire probes
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
z y
300 x
O
The raised floor
129.7
(b) 74
ϕ
3.3
96
1 : 3 Ahmed model
16.7
R3 z z
O O y
x
(c)
Roof
Rear window 40
7
Cp15
10
10
10
Cp22
10
5
5 Cp23
10
Cp9
Cp1 6.8
10
10
5
40
Rear 15
Vertical base
window
Vertical base
Figure 1. (a) Schematic of the experimental arrangement. (b) Dimensions of a one-third-scaled Ahmed body
(35°). (c) Pressure tap distribution on the rear end. The length unit is millimetre.
and x* ∈ [0.09, 1.52], y* ∈ [0, 0.72] and z* ∈ [−0.04, 0.90] behind the vertical base. The
total measurement points amount to more than 1500. The fast Fourier transform (FFT)
algorithm was employed to calculate the power spectral density function Eu of u. The FFT
window size Nw was 4096. The frequency resolution f in the spectral analysis depends
on the sampling frequency and the FFT window size (f = fs /Nw ) and is estimated to be
0.73 Hz (Zhou et al. 2012).
A Dantec high-speed two-dimensional PIV system (Litron LDY304-PIV, Nd:YLF) was
used at the HITsz wind tunnel to measure the wake of the Ahmed body. The model surface
and raised floor were painted black to minimize laser reflection noise. The flow was
seeded by smoke generated from peanut oil using a TSI 9307-6 particle generator, with
particles of about 1 μm in diameter. Flow illumination was provided by two pulsed laser
sources of a 527 nm wavelength, each with a maximum energy output of 30 mJ pulse−1 .
Particle images were taken using a charge-coupled device camera (PhantomV641, double
frames, with a resolution of 2560 × 1600 pixels). Synchronization between image taking
and flow illumination was provided by a Dantec timer box 80N77. PIV measurements
were performed in four x–z planes of y* = 0, 0.14, 0.36 and 0.54, and eight y–z planes,
i.e. x* = −1.27 at the roof and side surface of the body, x* = −0.36 and −0.09 above
the rear window, and x* = 0.20, 0.43, 0.65, 0.86 and 1.29 behind the base. The PIV
image covered an area of x* ∈ [−0.88, 2.80] and z* ∈ [−0.30, 1.99] for the x–z planes and
y* ∈ [−0.60, 0.98] and z* ∈ [−0.35, 2.17] for the y–z planes. The image magnifications in
both directions of each plane were identical, approximately 160 and 110 μm pixel−1 in the
x–z and y–z planes, respectively. The intervals between two successive pulses were 120
μs and 40 μs for measurements in the x–z and y–z planes, respectively. In processing PIV
images, the adaptive PIV method (Dynamic Studio software) was used with a minimum
interrogation area size of 32 × 32 pixels and a maximum size of 64 × 64 pixels. The
number of particle densities per interrogation window was at least 25 or 20 in the x–z or
y–z planes, respectively. The grid step size of 16 × 16 pixels produced 160 × 100 in-plane
velocity vectors and the same number of vorticity data in the x–z and y–z planes.
Following Zhang et al. (2018), the uncertainty of PIV measurements was estimated
based on image matching analysis. This method identifies particle image pairs in two
successive exposures according to the measured displacement vectors and evaluates the
residual distance or particle disparity vector between the particle image pairs, whose
ensemble within the interrogation window is used to infer the instantaneous uncertainty
of the measured displacement. For further details of this technique, please refer to
Sciacchitano, Wieneke & Scarano (2013). In the x–z planes, the estimated particle
displacement uncertainties are below 0.15 pixels in both x and z directions, resulting in
velocity uncertainties σ U and σ W , in U and W, less than 0.2 m s−1 or 2.5 % of U∞ ,
respectively. Following Wen, Tang & Duan (2015), the uncertainty σ ωy of the spanwise
vorticity ωy is given by
https://doi.org/10.1017/jfm.2020.1136
with a longer time period. The other is set at 500 Hz to achieve a higher temporal
resolution for instantaneous and proper orthogonal decomposition (POD) analyses. A
total of 2000 PIV images were captured in the double frame mode for each test run. The
percentage variations of all mean and root mean square velocities Ū ∗ , V̄ ∗ , W̄ ∗ , u∗rms , vrms
∗
∗
or wrms converge with increasing N to less than ±1 % once N > 1500, irrespective of the
measurement plane or trigger rate. As such, N = 2000 is considered to be adequate for
capturing the mean and fluctuating flow fields.
Laser-induced fluorescence (LIF) flow visualization was conducted at HKPU using a
Dantec standard PIV system (PIV2100). Please refer to Zhang et al. (2015) for the details
of the experimental set-up. Three rows of pinholes, each consisting of 64 circular orifices,
1 mm in diameter and equally separated by 1 mm, were made along lines parallel to and
7 mm upstream of the upper edge of the rear window, and the upper and lower edges of
the vertical base, respectively. Smoke generated by a TSI 3079 particle generator using
paraffin oil was pumped through one hollow cylindrical strut into a cavity in the rear part
of the model and released into the flow from the pinholes. Flow images were taken in
the x–z plane of y* = 0 and the x–y plane of z* = 0.72 over the rear window. The model
surface, the raised floor and the tunnel working section walls hit by the laser sheets were
all painted black to minimize reflection.
A PSI DTC Initium system was used to measure the time-averaged surface pressure
on the model at HITsz. Twenty-eight pressure taps were made, whose locations are
schematically shown in figure 1(c). The pressure taps were connected to an ESP scanner
via plastic tubes of 1 mm inner diameter. The scanner was placed inside the test model
to minimize the length of the tubes and hence the filtering effect of tubing in pressure
measurements (Grandemange, Gohlke & Cadot 2013). The measurement accuracy was
estimated to be approximately ± 2 Pa. The sampling frequency was 650 Hz, and a total
of 40 000 data points were obtained for each measurement location. The time-averaged
pressure coefficient is given by Cp = ( p − p0 )/(0.5ρU∞ 2 ), where p and p are the
0
time-averaged local and free-stream static pressures, respectively, and ρ is the density of
air.
The hot-wire and flow visualization measurements were carried out at U∞ ∈
[4, 36] m s−1 and 3 m s−1 , corresponding to Re ∈ [0.3, 2.7] × 105 and 0.22 × 105 ,
respectively. The PIV and pressure measurements were performed at U∞ = 8.3 m s−1
(Re = 0.62 × 105 ). The U∞ value was measured using a Pitot-static tube connected to an
electronic micro-manometer (Furness Control Ltd, FC510). Note that the Ahmed model
has sharp edges at the rear end, and the flow structure at this relatively low Re range does
not differ significantly from that at large Re of the order of 106 (Vino et al. 2005).
centre is tilted downstream from the symmetry plane to the off-symmetry plane. On the
other hand, the centre of the lower bubble shifts upstream, from (x*, z*) = (0.47, 0.07) to
(0.39, 0.09), from the plane of y* = 0.14 to that of y* = 0.36. In the x–z plane of y* = 0.54,
both bubbles disappear (figure 2d), and an upwash flow occurs over the rear window and
behind the base.
The vortex definition proposed by Zhou et al. (1999) is presently adopted for the
identification of vortices from the vorticity data and is briefly introduced below. A vortex
core is a region where the velocity gradient tensor ∇U has complex eigenvalues (Chong,
Perry & Cantwell 1990). Then, ∇U may be written as
⎡ ⎤
λr
∇U = [v r v cr v ci ] ⎣ λcr λci ⎦ [v r v cr v ci ]−1 , (3.1)
−λci −λcr
where λr is the real eigenvalue with a corresponding eigenvector v r , and λcr ± λci i are a
conjugate pair of complex eigenvalues with complex eigenvectors v cr ± v ci i. The local
flow is either stretched or compressed along the axis v r , while swirling in the plane
spanned by vectors v cr and v ci . The strength of the local swirling motion is quantified by
the imaginary part of the complex eigenvalue pair λci , referred to as the local swirling
strength of the vortex. This method is frame-independent and does not detect regions
containing significant vorticity but no local swirling motion, such as shear layers. A vortex
is identified if λ2∗
ci is larger than a threshold, 0.4 in the x–z planes and 0.002 in the y–z
planes, which are about 3 % and 1 % of the maximum λ2∗ ci , respectively.
As shown in the contours of instantaneous spanwise vorticity ωy∗ in each x–z plane from
y* = 0 to 0.54 (figure 3a,c,e,g), one highly concentrated negative (or positive) vorticity
layer separates from the upper edge of the rear window and another from the lower edge of
the base. The former in the plane of y* = 0.54 (figure 3g) is deflected downwards towards
the surface, as compared with its counterpart in the symmetry plane (figure 3a), spatially
coinciding with the well-known C-pillar vortex. The wake is characterized by two regions,
the upper of the negative ωy∗ concentrations and the lower of the positive, and the two
regions appear contracting from y* = 0 to 0.54. This contraction is confirmed by the ω̄y∗
contours (figure 3b,d, f,h) and is internally consistent with the streamlines (figure 2), where
both bubbles are even invisible in the plane of y* = 0.54 (figure 2d).
Interestingly, a small corner vortex with positive vorticity occurs at the lower end
of the rear window, which runs across almost the entire width of the window with
∗
ω̄y,max ≈ 4.0. In this paper, subscripts ‘max’ and ‘min’ denote extreme values in positively
and negatively signed vorticity concentrations, respectively. Wang et al. (2013) captured
a similar structure based on the PIV measurements in the x–z plane of y* = 0 behind an
Ahmed body (ϕ = 35°, Re = 0.62 × 105 ), and suggested that this structure was induced by
https://doi.org/10.1017/jfm.2020.1136
the negative circulation above the rear window. No further discussion was pursued. In the
present work, a different view on the formation of the corner vortex, as well as its effect on
the surface pressure, is provided based on the PIV, pressure and flow visualization data.
The time-averaged pressure coefficient Cp on the rear window (figure 4a) drops
gradually downstream from Cp1 = −0.32 at z* = 0.82 to Cp8 = −0.38 at z* = 0.50 in the
symmetry plane of y* = 0. On the other hand, Cp on the vertical base increases almost
linearly from Cp9 = −0.27 at z* = 0.39 to Cp14 = −0.18 at z* = 0.13. Similar observations
913 A21-8
Flow structure around a low-drag Ahmed body
(a)
1.2
1.0
0.8
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
z∗ 0.6
0.4
y∗ = 0
0.2
(b)
1.2
1.0
0.8
z ∗ 0.6
0.4
y ∗ = 0.14
0.2
(c)
1.2
1.0
0.8
z∗ 0.6
0.4
y ∗ = 0.36
0.2
(d )
1.2
1.0
0.8
z∗ 0.6
0.4
y ∗ = 0.54
0.2
0
–0.5 0 0.5 1.0 1.5
x∗
https://doi.org/10.1017/jfm.2020.1136
Figure 2. PIV-measured time-averaged sectional streamlines in x–z planes of (a) y* = 0, (b) y* = 0.14, (c)
y* = 0.36 and (d) y* = 0.54 at Re = 0.62 × 105 . The red ‘+’ denotes the focus of streamlines.
913 A21-9
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
0.8
z ∗ 0.6
0.4
y∗ = 0 y∗ = 0 12.3
0.2
0
(c) (d )
1.2
Corner –9.3
1.0 vortex 3.9
0.8
z ∗ 0.6
0.4
y ∗ = 0.14 y ∗ = 0.14 11.5
0.2
0
(e) (f)
1.2 Corner
vortex –16.2 4.2
1.0
0.8
z ∗ 0.6
0.4
y ∗ = 0.36 y ∗ = 0.36 12.7
0.2
0
(g) (h)
1.2
Corner –14.2
1.0 vortex 4.0
0.8
z ∗ 0.6
0.4
y ∗ = 0.54 y ∗ = 0.54 –4.5
0.2 9.0
0
–0.5 0 0.5 1.0 1.5 –0.5 0 0.5 1.0 1.5
x∗ x∗
Figure 3. PIV-measured typical instantaneous (a,c,e,g) and time-averaged (b,d, f,h) spanwise vorticity
contours in x–z planes of (a,b) y* = 0, (c,d) y* = 0.14, (e, f ) y* = 0.36 and (g,h) y* = 0.54 at Re = 0.62 × 105 .
Black thick lines are contours of the swirling strength λ̄2∗
ci = 0.4 of the time-averaged flow. The magnitudes of
∗
ω̄y,max ∗
and ω̄y,min are marked in the figure.
https://doi.org/10.1017/jfm.2020.1136
have been made from Cp15 to Cp28 in the off-symmetry plane of y* = 0.36. The Cp
distribution is in reasonable agreement with previous measurements on the rear end of
an Ahmed body with ϕ = 35° by Lienhart & Becker (2003) and Meile et al. (2016). There
is an appreciable deviation in Cp , which is ascribed to the differences in experimental
conditions such as Re and facilities among these studies. The pressure on the vertical
913 A21-10
Flow structure around a low-drag Ahmed body
(a) (b)
0.8
Corner
vortex
0.7
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
0.6
0.2
0.1
0
–0.4 –0.3 –0.2 –0.1
Cp
Figure 4. (a) Time-averaged pressure coefficients Cp on the rear end in the symmetry plane of y* = 0 (red ‘×’)
and an off-symmetry plane of y* = 0.36 (black ‘◦’) at Re = 0.62 × 105 . The blue ‘’ and green ‘∇’ symbols
denote Cp experimentally obtained in the plane of y* = 0 by Lienhart & Becker (2003) (Re = 8.93 × 105 ) and
Meile et al. (2016) (Re = 6.70 × 105 ), respectively, for an Ahmed body with ϕ = 35°. (b) A typical photograph
of the flow structure in the symmetry plane at Re = 0.22 × 105 . Flow is from left to right.
base is appreciably higher than that on the rear window. The flow within the upper
recirculation bubble near the vertical base moves upwards along the base under the
pressure gradient, which is evident in figure 3. Such upward-moving fluid separates at
the upper edge of the base and then rolls up before reattachment on the slanted surface
under the pressure difference between the two surfaces, forming the corner vortex, which
is very well illustrated by one typical photograph captured from flow visualization in the
symmetry plane (figure 4b).
Grandemange et al. (2013) observed a bistability in the wake of a square-back Ahmed
body, which is characterized by random spanwise switchings of the recirculation region
between two preferred reflectional symmetry-breaking positions. The same observation
was also made and investigated by Östh et al. (2014) and Barros et al. (2017). This
phenomenon has, however, never been observed in the wake of the low-drag regime.
The measured time-averaged drag coefficient CD (≡FD /(0.5ρU∞ 2 A), where F is
D
aerodynamic drag) is 0.32, slightly higher than previous reports by Ahmed et al. (1984),
Guilmineau (2008), Meile et al. (2016) and Rao et al. (2018), which is probably due to
the present lower Re. It is evident in table 2 that CD decreases from 0.32 to 0.26 from
Re = 1.67 × 105 to 1.4 × 106 .
https://doi.org/10.1017/jfm.2020.1136
Researchers Re CD
Ahmed et al. (1984) 1.40 × 106 0.26
Guilmineau (2008) 8.93 × 105 0.27
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
increases to 2.1 downstream in the plane of x* = −0.09 (figure 5d). This longitudinal
structure is apparently the C-pillar vortex, which is formed due to the rollup of the shear
layer coming off the sidewall about the side edge of the rear window because of the
pressure difference between the flows over the sidewall and the slanted surface (Ahmed
et al. 1984). The strength of the C-pillar vortex is substantially lower than that in the
∗
high-drag regime; its ω̄x,max in the y–z plane of x* = 0.20 was only about 10 % of the
∗
latter (Wang et al. 2013). It is noteworthy that the ω̄x,max location of the C-pillar vortex
moves inwards and downwards from (y*, z*) = (0.59, 0.82) at x* = −0.36 to (0.53, 0.70)
at x* = −0.09. Its angles are about 11° and 12° with respect to the slanted surface and the
sidewall, respectively, larger than those (8° and 7°) at ϕ = 25° (Lehugeur et al. 2005).
As shown in the ω̄x∗ contours in the y–z plane of x* = −0.09 (figure 5d), a relatively
strong ω̄x∗ concentration is present in the vicinity of the C-pillar vortex, whose sign
is opposite to that of the C-pillar vortex, with ω̄x,min∗ reaching −0.8, though without
swirling. This vorticity concentration is induced by the C-pillar vortex (Zhang et al. 2015).
A number of alternately signed ωx∗ concentrations are discernible at the same height as
the roof in the y–z planes of x* = −0.36 and −0.09 (figure 5a,c), which are probably
associated with the hairpin vortices wrapped with the longitudinal vortices above the roof.
There is another streamwise swirling structure enclosed by the contour of λ̄2∗ ci = 0.002
near the lower corner of the sidewall in both planes. This structure, referred to as the lower
vortex, is formed along the lower edge of the sidewall (Wang et al. 2013).
It is insightful to examine the ω̄x∗ contours, superimposed with sectional streamlines,
in different y–z planes behind the vertical base (figure 6). At x* = 0.20 (figure 6a),
the upper rightmost concentration marked by ‘C’ may be identified with the C-pillar
vortex. Apparently, the C-pillar vortex starts to decay from ω̄x,max ∗ =2.0 to 1.3 from
x* = 0.20 to 0.65 (figure 6a–c). A longitudinal swirling structure (marked by the contour
of λ̄2∗ ∗
ci = 0.002) with ω̄x,min =−1.7 occurs in x* = 0.20 on the inner side, close to the
symmetry plane, of the C-pillar vortex, probably resulting from the interaction between the
C-pillar vortex and the upper recirculation bubble. This induced structure starts to decay
downstream with the C-pillar vortex. Behind the base, there is a large swirling structure
enclosed by the contour of λ̄2∗ ci = 0.002 (figure 6a), which coincides spatially with the
upper bubble. Its vorticity concentration increases from ω̄x,max ∗ =2.2 at (y*, z*) = (0.45,
∗
0.58) in the y–z plane of x* = 0.20 to ω̄x,max =3.1 at (y*, z*) = (0.38, 0.52) in the plane
https://doi.org/10.1017/jfm.2020.1136
∗
of x* = 0.43 (figure 6a,b). In both planes, the ω̄x,max value occurs at almost the same
height (z* ≈ 0.52) as the upper bubble centre (figure 2). When moving downstream to
∗
x* = 0.65, ω̄x,max drops to 2.6 (figure 6c). Based on the streamlines in the x–z planes of
y* ∈ [0, 0.36] (figure 2a–c), this longitudinal structure is linked to the upper bubble, which
runs laterally outwards downstream. This longitudinally tilted large structure induces an
upwash flow on its outer side (figure 6a–c), responsible for the upwash captured in the
913 A21-12
Flow structure around a low-drag Ahmed body
x ∗ = –0.36 x ∗ = –0.09
(a) (c)
1.2
C-pillar C-pillar ω∗x
1.0 vortex vortex
7.2
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
0.8 4.8
z ∗ 0.6 2.4
–0.8
0.4
–3.2
0.2
–5.6
0 –8.0
(b) (d )
1.2 –0.4 –0.3 ω̄∗x
1.0 1.8 –0.8 2.0
2.1
1.6
0.8 1.2
z ∗ 0.6 0.8
0.4
0.4 –0.2
1.5 1.4
0.2 –0.6
–1.0
0 –1.4
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
y∗ y∗
Figure 5. PIV-measured instantaneous (a,c) and time-averaged (b,d) streamwise vorticity contours in the y–z
planes of x* = −0.36 (a,b) and x* = −0.09 (c,d) at Re = 0.62 × 105 . The thick black curves are the contours of
the swirling strength λ̄2∗ ∗ ∗
ci = 0.002 of the time-averaged flow. The magnitudes of ω̄x,max and ω̄x,min are marked
in the figure.
x–z plane of y* = 0.54 (figure 2d), and a downwash flow on the inner side close to the
symmetry plane (figure 6b,c). Both the C-pillar vortex and the large longitudinal structure
move downwards and towards the symmetry plane from x* = 0.43 to 0.65 (figure 6b,c)
under a pressure difference between the outer flow and the wake. From x* = 0.86, the
C-pillar vortex appears to merge into the longitudinal large structure (figure 6d,e), that is,
the far wake is dominated by one single pair of longitudinal trailing vortices. The lower
∗
vortex (figure 5) decays gradually in the wake from ω̄x,max ∗
=1.5 at x* = 0.20 to ω̄x,max =0.8
at x* = 1.29 (figure 6).
rather pronounced peak occurs at f* = Stw = 0.30 at B1 (z* = 0) and B3 (z* = 0.25) at
x* = 0.43, C1 (z* = 0), C3 (z* = 0.25) and C4 (z* = 0.32) at x* = 0.65, and D1 (z* = 0)
and D3 (z* = 0.25) at x* = 0.86, but not at B2 , C2 and D2 (z* = 0.13), where a minor
peak is discernible only at the second harmonic of Stw = 0.30, i.e. f* = 0.60. The peak at
Stw = 0.30 becomes more pronounced with increasing x* for lower positions B1 through
D1 and more visible at higher z* (B3 , z* = 0.25; C4 , z* = 0.32) from x* = 0.43 to 0.65.
913 A21-13
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
(a) x ∗ = 0.20
1.2
1.0 –1.7 C (b) x ∗ = 0.43 ω̄∗x –3.6 –2.7 –1.8 –0.9 0.3 1.2 2.1 3.0
2.0
0.8
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
A6
A5 z∗
0.72
A4 B6
A3 0.49 B5 z∗
0.54
Eu A2 0.40 B4 C6
z 0.39 z∗
ϕ = 35° y 0.27 C5
10–2 A B3 0.54
1 0.32
0.13 D6
o 10–3 B2 C4
10–4 –2 0 0.25 0.39
z∗
10 10–1 100 f∗ B1
C3 0.32
D5
0.54
0.13 0.60 D4 E6
A1–A6 Stw = 0.30
(x∗ = 0.20, y∗ = 0) 0
C2 0.25 0.39 z∗
10–2 10–1 100 D3 E5 0.54
C1 0.13 0.32
B1–B6 Stw = 0.30 E4 0.39
D2
(x∗ = 0.43, y∗ = 0) 0 0.25
10–2 10–1 100 E3
D1 0.32
C1–C6 Stw = 0.30 0.13
E2
(x∗ = 0.65, y∗ = 0) 0 Sts = 0.27
0.25
10–2 10–1 100
E1
0.13
D 1– D 6
https://doi.org/10.1017/jfm.2020.1136
(x∗ = 0.86, y∗ = 0) 0
10–2 10–1 100
E1–E6
(x∗ = 1.29, y∗ = 0) x
Figure 7. Power spectral density function Eu of the hot-wire signal measured along A1 –A6 , B1 –B6 , C1 –C6 ,
D1 –D6 and E1 –E6 , respectively, at y* = 0 in the wake, at Re = 0.62 × 105 .
913 A21-14
Flow structure around a low-drag Ahmed body
The observations suggest that the unsteady structures associated with Stw = 0.30 are
related to the two recirculation bubbles behind the base. The peak at Stw = 0.30 is no
longer visible for x* ≥ 1.29, suggesting the disappearance of these unsteady structures.
Zhang et al. (2015) captured a predominant frequency of Stw = 0.44, larger than the
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
present Stw , behind the vertical base of an Ahmed model with ϕ = 25° (Re = 0.62 × 105 ).
Other investigators (e.g. Minguez et al. 2008; Joseph, Amandolese & Aider 2012; Kohri
et al. 2014) also reported this predominant frequency, varying rather wildly, in the wake of
an Ahmed body with various ϕ (figure 8a). One may wonder whether there is any scaling
law to govern this predominant frequency. In the high-drag regime, the flow separates from
both the upper and lower edges of the base, forming two recirculation bubbles behind the
base. The Stw is attributed to the alternate emanation of unsteady structures from the two
bubbles (Zhang et al. 2015). We may surmise that the height h of the vertical base is a
characteristic length in this case. On the other hand, H might be the characteristic length
in the low-drag regime where the two recirculation bubbles result from flow separation
from the upper edge of the rear window and the lower edge of the base. Note that the
Ahmed body with the critical angle ϕ = 30° can be in either high- or low-drag regimes
(Sims-Williams 2001; Conan, Anthoine & Planquart 2011; Kohri et al. 2014). Furthermore,
it was observed even in the high-drag regime that the separated flow over the rear window
did not fully reattach on the slanted surface and, instead, merged with the upper bubble
behind the base (Vino et al. 2005; Conan et al. 2011). That is, H should be considered as
the characteristic length for ϕ = 30° in both regimes. As such, we may define
h, ϕ < 30◦ ,
Stw# = fS/U∞ , S = (3.2a,b)
H, ϕ ≥ 30◦ ,
where S is the vertical distance between the separation points of the two bubbles. The
square-back Ahmed body may be described by ϕ = 0° or 90°, corresponding to S = h or
H, respectively, where h equals H.
The dependence of Stw# on ϕ is presented in figure 8(b). The thus defined Stw#
displays a good collapse, increasing from 0.23 to about 0.30 as ϕ varies from 12.5°
to 30°, and dropping once ϕ > 30°. It seems plausible that Stw indeed scales with S.
There is an appreciable variation in Stw# , especially for ϕ < 30°. Zhang et al. (2015)
discovered at ϕ = 25° that the unsteady structures of Stw resulted from the nearly periodic
growth-and-burst cycle of the two bubbles behind the base. Naturally, Stw may depend on
the bubble size, and a big bubble is associated with a long growth period before bursting
and thus a low Stw . On another note, the bubble size is affected by the effective separation
angle θ , with respect to the streamwise direction, of the separated shear layer. A large θ
may act to cut short the bubble size, thus increasing Stw# . For the high-drag regime, θ is
obviously the same as ϕ; for the low-drag regime, flow separates from the upper edge of
the rear window and θ should be given by the angle between the streamwise direction
and the local tangential velocity at x* = 0 of the streamline through the separation point
(figure 8c). The choice of the point at x* = 0 on the streamline ensures that the effective
separation angles are taken at the same streamwise location. For the square-back Ahmed
body, θ is 0°. Then, we may define a new non-dimensional frequency by
https://doi.org/10.1017/jfm.2020.1136
90◦ − θ
Stw+ = αStw# = f αS/U∞ , α= , (3.3a,b)
90◦
where α is a deflection coefficient with a value of α ∈ [0, 1], depending on θ. This Stw+
collapses well to a constant of about 0.20, regardless of ϕ (figure 8c). For the present data,
θ is estimated to be 13° from the PIV-measured time-averaged streamlines in figure 2(a).
913 A21-15
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
(a) 0.7
0.6
0.5
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
0.4
Stw
0.3
0.2
0.2 0.2
h, ϕ < 30°
0.1
{
St#w = f S/U∞, S = H, ϕ ≥ 30° 0.1 0.20
0 0
10 20 30 90 10 20 30 90
ϕ (deg.) ϕ (deg.)
Figure 8. Dependences of (a) Stw , (b) Stw# and (c) Stw+ on ϕ. Symbols: ◦, present measurement
(Re = 0.62 × 105 ); ×, Kohri et al. (2014, Re = 0.73 × 105 ); +, Zhang et al. (2015, Re = 0.89 × 105 ); ,
Sims-Williams (2001, Re ∈ [3.10, 6.40] × 105 ); ∇, Vino et al. (2005, Re = 4.40 × 105 ); , Grandemange
et al. (2013, Re = 1.07 × 105 ); , Barros et al. (2016a, Re = (2.33–6.98) × 105 ); , Joseph et al. (2012,
Re = 4.49 × 105 ); ♦, Minguez et al. (2008, Re = 8.93 × 105 ); ✩, Volpe, Devinant & Kourta (2015,
Re = 5.42 × 105 ); , Lahaye, Leroy & Kourta (2014, Re = 6.41 × 105 ). The uncertainty bars for the present
data are determined from the standard deviation of six repeated measurements.
We use θ = ϕ = 30° when rescaling Sims-Williams’ (2001) and Vino et al.’s (2005) data
in the high-drag regime. Kohri et al.’s (2014) data at ϕ = 30°, but corresponding to the
low-drag regime, are not included in figure 8(c), as the velocity field is not reported and
hence θ could not be estimated. The slight scatter in the Stw+ data is probably due to a
difference in facilities, Re and measurement uncertainties. The product of αS in (3.3a,b)
may be interpreted as a corrected characteristic length of the separation bubbles, which
depends on both S and θ.
https://doi.org/10.1017/jfm.2020.1136
It is of interest to understand the physics behind the unsteady structures of Stw = 0.30.
Figure 9 shows PIV-measured instantaneous flow structures in one typical cycle with a
period Tw∗ (≈3.3) of the unsteady structure generation in the symmetry plane of the wake.
Following Zhang et al. (2015), the length l* of each bubble is defined by the maximum
longitudinal length of the region where U ≤ 0. At the beginning (t* = 0) of the cycle,
one organized structure characterized by the concentration of quite strong positive ωy∗
913 A21-16
Flow structure around a low-drag Ahmed body
1.0
0.8
z ∗ 0.6
0.4
t∗ = 0 t ∗ = 0.25T ∗w
0.2
0
(c) (d )
1.2
1.0
0.8
z ∗ 0.6
0.4
t ∗ = 0.5T ∗w t ∗ = 0.75T ∗w
0.2
0
–0.5 0 0.5 1.0 1.5 –0.5 0 0.5 1.0 1.5
x∗ x∗
Figure 9. PIV-measured instantaneous sectional streamlines and isocontours of spanwise vorticity ωy∗ in one
typical cycle in the symmetry plane of y* = 0 at Re = 0.62 × 105 . The red solid curve indicates the bubble size,
determined from the sectional streamline that encloses the region of recirculation bubbles. The red broken line
highlights the structure emanated from each bubble.
(figure 9a), as highlighted by a red broken curve, appears pinched off from the lower
recirculation bubble at x* ≈ 0.4, shooting slightly upwards. At t* = 0.25Tw∗ , the upper
recirculation bubble undergoes a decrease in size from l* = 0.76 at t* = 0 to l* = 0.60,
accompanied by the emanation of a structure of negative ωy∗ from it (figure 9b). Then,
both bubbles gradually grow in size, with l* increasing to 0.81 at t* = 0.75Tw∗ (figure 9c,d).
During the whole cycle, the upper and lower bubbles undergo a nearly periodic expansion
and contraction, l* varying over [0.60, 0.81] and [0.33, 0.81], respectively. Kohri et al.
(2014) detected a predominant frequency St ≈ 0.27 behind the Ahmed body of ϕ = 35°
at Re = 0.73 × 105 , which was ascribed to the fluctuation of an upwash flow from the
lower edge of the base. They advocated that the fluctuation was induced by periodic
vortex shedding from the bottom end. However, the present hot-wire and PIV data suggest
a different scenario. The unsteady structures of Stw = 0.30 may result from the nearly
periodic expansion and contraction of both bubbles when the organized structures are
either pinched off from the lower bubble or emanated from the upper bubble, with their
onset at x* ≈ 0.4. More elaboration on these unsteady structures and nearly periodic
variation of the bubbles will be given below.
https://doi.org/10.1017/jfm.2020.1136
The spectral coherence Cohu1 u2 (= (Co2u1 u2 + Q2u1 u2 )/Eu1 Eu2 ) between two fluctuations
u1 and u2 provides a measure for the correlation between the spectral components of u1
and u2 (Antonia, Zhou & Matsumara 1993), where Cou1 u2 and Qu1 u2 are the co-spectrum
and quadrature spectrum of u1 and u2 , respectively. The spectral coherence CohuB1 uB3
(figure 10a) between two simultaneously measured hot-wire signals at B1 and B3
(x* = 0.43) is characterized by one predominant peak around Stw = 0.30, reaching nearly
913 A21-17
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
(a) 0.4 (b) π
f ∗ = 0.30
0.3 0.5π
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
1 uB3
1 uB3
f ∗ = 0.30
CohuB
0.2 0
ΦuB
0.1 –0.5π
0 –π
10–2 10–1 100 101 0 0.1 0.2 0.3 0.4 0.5 0.6
f∗ f∗
B3
ϕ = 35° y
o B1 x
Figure 10. (a) Spectral coherence CohuB1 uB3 and (b) phase ΦuB1 uB3 between simultaneously measured
hot-wire signals uB1 and uB3 , at Re = 0.62 × 105 . Please refer to figure 7 for the locations of B1 and B3 .
0.35, and the corresponding spectral phase shift ΦuB1 uB3 (= tan−1 (QuB1 uB3 /CouB1 uB3 )) is
about 0.5π (figure 10b). B1 and B3 (and also C1 and C3 or D1 and D3 that follow) are
as in figure 7. Similar observations are also made at x* = 0.65 (C1 and C3 ) and x* = 0.86
(D1 and D3 ), corresponding to a phase shift of 0.63π and 0.70π (not shown), respectively.
Note that there is a difference in the convecting velocity between the flows at z* = 0 and
0.25, that is, the velocity of the flow from the gap between the body and the ground
is higher than that in the wake around z* = 0.25 (not shown). This difference varies
downstream in the near wake, accounting for the increase in Φ from x* = 0.43 to 0.86.
Clearly, the unsteady structures detected at B1 and B3 , C1 and C3 or D1 and D3 are highly
correlated. This phase shift is distinct from that of near π associated with the unsteady
structures emanated from the two recirculation bubbles behind the base in the high-drag
regime (Zhang et al. 2015).
To gain an in-depth understanding of the unsteady structures of Stw = 0.30 and why
ΦuB1 uB3 ≈ 0.5π, we performed a POD analysis of the time-resolved PIV data obtained
over x* ∈ [−0.61, 1.93] and z* ∈ [−0.07, 1.34] in the symmetry plane. The POD is
introduced in detail by Sirovich (1987), for example. Briefly, the spatiotemporal fluctuating
velocity field u(x, t) is decomposed into an optimal set of spatial base functions Ψ n (x) and
temporal mode coefficients an (t), viz.
u(x, t) = U(x, t) − Ū(x) = an (t)Ψ n (x), (3.4)
https://doi.org/10.1017/jfm.2020.1136
where U is the instantaneous velocity vector at the position x and n is the POD mode
number. The fluctuating velocity fields from M snapshots can be arranged in a matrix uM
as
uM = [ u(t1 ) u(t2 ) ... u(tM ) ]. (3.5)
913 A21-18
Flow structure around a low-drag Ahmed body
(a) (b) 0.08
0.12 Mode 1
Mode 2
0.10 Mode 3 0.06
Mode 4
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
0.08
λn / (λn)
0.04
0.06
0.04
0.02
0.02
0
400 600 800 1000 0 5 10 15 20
M n
Figure 11. (a) POD eigenvalue convergence as a function of PIV snapshot number M. (b) Relative contribution
of individual POD modes for mode n ∈ [1, 20], where M = 1000. The POD analysis is conducted over
x* ∈ [−0.61, 1.93] and z* ∈ [−0.07, 1.34] in the symmetry plane. Here Re = 0.62 × 105 .
100 100
Power spectrum
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
80 80
Mode 1 Mode 2
60 60
40 40
20 20
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0
(c) (d )
140 140
120 120
100 100
Power spectrum
80 80
Mode 3 Mode 4
60 60
40 40
20 20
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0
f∗ f∗
Figure 12. Power spectra of the first four POD coefficients: (a) mode 1; (b) 2; (c) 3; and (d) 4. The POD
analysis is conducted over x* ∈ [−0.61, 1.93] and z* ∈ [−0.07, 1.34] in the symmetry plane and M = 1000. Here
Re = 0.62 × 105 .
ũ(x, φ(t)), obtained by a POD reconstruction. The power spectral density functions of the
POD coefficients display a pronounced peak at f* ≈ 0.30 for modes 1 and 2 (figure 12a,b)
but not for modes 3 and 4 (figure 12c,d) or higher modes (not shown). The observation
suggests that only modes 1 and 2 are associated with the nearly periodic structures of
Stw = 0.30.
A phase-resolved flow field may be reconstructed approximately with a low-order flow
model including the mean flow (mode 0) and modes 1 and 2 (Ben Chiekh et al. 2004), viz.
U LOM (x, φ) = Ū(x) + a1 (φ)Ψ 1 (x) + a2 (φ)Ψ 2 (x), (3.10)
√ √
where a1 = 2λ1 sin(φ) and a2 = 2λ2 cos(φ) (van Oudheusden et al. 2005), viz.
a1 (i) a2 (i)
√ = ri sin(φi ), √ = ri cos(φi ).
https://doi.org/10.1017/jfm.2020.1136
(3.11a,b)
2λ1 2λ2
The ri in (3.11) is the ‘radius’ of the normalized ellipse, viz.
a1 (i)2 a2 (i)2
ri2 = + . (3.12)
2λ1 2λ2
913 A21-20
Flow structure around a low-drag Ahmed body
1.0
0.8
z ∗ 0.6
0.4 10.0 –1.7
φ = 0° φ = 180°
0.2 6.1
0
(b) (f)
1.2
1.0
0.8
z ∗ 0.6
0.4 –1.6
φ = 45° 9.6 φ = 225°
0.2 4.3
0
(c) (g)
1.2
1.0
0.8
z ∗ 0.6
0.4 –2.3
φ = 90° 9.2 φ = 270°
0.2 2.1
0
(d ) (h)
1.2
1.0
0.8
z ∗ 0.6
0.4 –2.1
10.0
φ = 135° 7.8 φ = 315°
0.2 1.9
0
–0.5 0 0.5 1.0 1.5 –0.5 0 0.5 1.0 1.5
x∗ x∗
Figure 13. Phase-resolved reconstructed sectional streamlines (white) and spanwise vorticity within one
typical cycle in the symmetry plane y* = 0. The white symbol ‘×’ denotes the saddle point. The red line
indicates the bubble size, determined from the sectional streamline that encloses the region of recirculation
bubbles. Black thick lines are contours of λ2∗ ∗ ∗
ci = 0.4. The magnitudes of ωy,max and ωy,min are marked in the
figure. Here Re = 0.62 × 105 .
https://doi.org/10.1017/jfm.2020.1136
φ = 0° φ = 90°
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
(c) (d)
φ = 180° φ = 270°
Figure 14. Schematic of the physical process for the nearly periodic motion of two recirculation bubbles and
separated flow structures that result.
point B1 (x* = 0.43, z* = 0), appears to pinch off from the lower recirculation bubble
(figure 13a). The observation conforms to the emanated structure from the lower bubble,
as captured in figure 9(a). As a result, the length of the lower bubble retreats markedly
from its maximum l* = 0.82 at φ = 315° (figure 13h) to l* = 0.30 at φ = 0° (figure 13a).
Then, the lower recirculation bubble grows gradually in size from φ = 0° to 315°, because
of the continuous entrainment of fluid from the lower edge, until part of the bubble is
pinched off again in the next cycle. Obviously, the lower bubble undergoes a nearly
periodic variation in size over l* ∈ [0.30, 0.82]. This is internally consistent with the
observation from the instantaneous flow structure images in figure 9. Meanwhile, the
separated structure is advected downstream and its maximum vorticity decays from
∗
ωy,max ∗
=10.0 at x* = 0.66 (φ = 0°) to ωy,max =1.9 at x* = 1.16 (φ = 315°) (figure 13a–h).
This structure is invisible beyond x* = 1.5 during the entire cycle, implying that this
structure vanishes rather rapidly.
The upper recirculation bubble also varies in size. At φ = 45°, this bubble grows
to its maximum with l* = 0.90 (figure 13b). With continuously entrained fluid from
the upper edge of the slanted surface, the bubble cannot maintain itself any more and
bursts at φ = 90°, causing its size to collapse from l* = 0.90 at φ = 45° to l* = 0.69
at φ = 90° (figure 13b,c). Then, the bubble grows again, reaching l* = 0.81 at φ = 315°
(figure 13c–h). This growth-and-burst process reflects the upper bubble variation captured
in figure 9. The burst is accompanied by the separation of one swirling structure, marked
by the contour of λ2∗ ∗
ci = 0.4 with ωy,min =−2.3, from the upper recirculation bubble. This
structure is then advected downstream (figure 13c–f ). The result is consistent with the
structure emanation captured in figure 9(b). This structure decays quickly downstream,
∗
from ωy,min ∗
=−2.3 at φ = 90° to ωy,min =−1.6 at φ = 225°, and vanishes before x* = 1.0
at φ = 270° (figure 13c–g). Note that the emanation of structures from the lower and
upper bubbles occurs at φ = 0° and 90°, respectively, which agrees approximately with
https://doi.org/10.1017/jfm.2020.1136
than the lower one (figure 13). This difference is markedly bigger than that in the high-drag
regime (Zhang et al. 2015). The sudden contraction and pressure drop of the lower bubble
may naturally disturb the upper bubble and induce its bursting at φ = 90°, accompanied by
the emanation of a coherent structure. This may explain the phase shift of about 0.5π at
f* = 0.30 (figure 10b). This phase shift may imply a relatively strong interaction between
the structures emanated from the two bubbles and hence a short lifespan, vanishing by
x* = 1.29 (figure 7). In contrast, their counterparts in the high-drag regime are in antiphase
and persist till x* = 1.40 (Zhang et al. 2015). After bursting, the bubble starts its growth in
size and pressure of the next cycle, and the separated structures are advected downstream,
though decaying rapidly.
The second moments of the fluctuating velocities in the wake may have an important
impact on the surface pressure on the rear part of a bluff body (e.g. Balachandar, Mittal &
Najjar 1997; Barros et al. 2016b). Following Grandemange et al. (2013), we may write the
streamwise mean momentum balance of the recirculation region, where the viscous force
is neglected, behind a low-drag Ahmed body as
+2 u∗ w∗ (n · iz ) dB∗ , (3.13)
∂B
where ix , iy and iz are the unit vectors of the coordinate system, n is the local unit vector
normal to the differential element dB* of the three-dimensional recirculation boundary
∂B, Cp is the spatially averaged Cp on the rear end, and Cpw is the local time-averaged
pressure coefficient in the wake. Equation (3.13) shows a close connection between the
Reynolds stresses along the recirculation region boundary and the pressure drag. Given
the symmetry conditions in the plane of y* = 0, n · iy is zero so that the term u∗ v ∗ (n · iy )
vanishes in the mean recirculation region, as Barros et al. (2016b) did for a square-back
Ahmed body. Then, the sum of the integrations on the right-hand side of (3.13) may be
reduced to a two-dimensional form (Balachandar et al. 1997), namely,
occurs behind the rear window, which is apparently associated with the shear layer
separated from the upper edge of the slanted surface. The other takes place downstream of
the lower edge of the base. In the symmetry plane, this concentration spatially coincides
with the structure of Stw = 0.30 detected behind the base (figure 7) and is connected to the
nearly periodic motions of the upper and lower bubbles and hence the structures emanated
from the bubbles. Secondly, the w∗ w∗ contours display one highly concentrated region
913 A21-23
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
u∗u∗ 0.002 0.016 0.030 0.044 0.058 w∗w∗ 0.002 0.012 0.022 0.032 0.042 u∗w∗ –0.016 –0.004 0.010 0.022 0.034
behind the vertical base (figure 15c,d). The rollup of the gap flow and the downwash flow
from the rear window appear to clash with each other near (x*, z*) ≈ (0.78, 0.08) in the
symmetry plane of y* = 0 and (0.75, 0.42) in the off-symmetry plane of y* = 0.36 based
on the PIV-measured streamlines (figure 2a,c), which may account for this concentration.
Thirdly, two opposite-signed concentrations are evident in the u∗ w∗ contours in both x–z
planes (figure 15e, f ). The upper is negative while the lower is positive, which are linked
to the upper and lower recirculation bubbles, respectively.
In order to evaluate the effect of the u∗ u∗ and u∗ w∗ concentrations of each bubble
on Cp , ∂b in the symmetry plane is divided into two parts (∂bu and ∂bl ) by the
dividing point where u∗ w∗ equals zero, as marked by the black symbol ‘×’ at (x*,
z*) ≈ (0.78, 0.2) (figure 15a,g). The subscripts u and l denote the upper and lower bubbles,
respectively. Then, the sum of the integrations of u∗ u∗ and u∗ w∗ along ∂bu and ∂bl can be
written as
The CuR and ClR are calculated to be approximately −0.005 and −0.007 in the symmetry
plane, which contribute negatively to Cp . As will be shown in § 4, the structures of
Stw = 0.30 show up in the region of x* ∈ [0.43, 0.86], y* ∈ [−0.29, 0.29] and z* ∈ [−0.04,
0.32]. The Reynolds stress distribution in an off-symmetry plane of y* = 0.14 also displays
the concentrations of u∗ u∗ max = 0.052 at (x*, z*) = (0.69, 0.02) and u∗ w∗ max = 0.034
at (x*, z*) = (0.81, 0.05) (not shown), due to the presence of this unsteady structure. It
may be inferred from the relationship between the Reynolds stresses and Cp that this
https://doi.org/10.1017/jfm.2020.1136
unsteady structure may have a considerable impact on the base and rear window pressures
in the off-symmetry planes, as in the symmetry plane. The structures emanated alternately
from the upper and lower bubbles may produce a substantial turbulent kinetic energy
k* = 0.5(u∗ u∗ + w∗ w∗ ), as supported by the u∗ u∗ and w∗ w∗ concentrations associated
with these structures (figure 15). If we consider a vehicle cruising in ambient still fluid,
the energy loss due to the turbulent kinetic energy consumes part of the work done on
913 A21-24
Flow structure around a low-drag Ahmed body
(a) Str = 0.20
(b) Sts = 0.27
z∗
y∗ 0.86 G7
0.43 F5
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
0.72 G6
0.36 F4 0.57 G5
0.43 G4
Sta = 0.14 0.29 F3
0.29 G3
10–2
0.14 F2 10–2
Eu 10–3 Str = 0.20 0.14 G2
10–3 St = 2.12
0 F1 l 0 G1
10–4 –2 10–4 –2
10 10–1 100 10 10–1 100
f∗ f∗
F1
z F5
G7
x y
o
G1
(c) (d )
1.2 ω∗x 1.2 –0.8 ω̄∗x
10 2.0
1.0 1.0
1.6
6
0.8 0.8 0.8 1.2
z∗ 0.6 2 0.6 0.8
–3 0.4
0.4 0.4
–0.2
0.2 –7 0.2 –0.3
2.2 –0.6
0 –11 0 –1.0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
y∗ y∗
Figure 16. Power spectral density function Eu of the hot-wire signal measured along (a) F1 –F5 (x* = −1.27,
z* = 0.94) above the roof, and (b) G1 –G7 (x* = −1.27, y* = 0.63) near the sidewall, respectively. PIV-measured
(c) typical instantaneous and (d) time-averaged streamwise vorticity contours in the y–z plane of x* = −1.27.
Thick black closed contours correspond to the swirling strength λ̄2∗ ci = 0.002 of the time-averaged flow. The
∗
magnitudes of ω̄x,max ∗
and ω̄x,min are marked in the figure. Here Re = 0.62 × 105 .
the fluid, and hence increases the vehicle pressure drag. That is, the emanated structures
contribute to a rise in the drag.
instantaneously above the roof around y* = 0.14 (figure 16c), which is responsible for the
time-averaged swirling structure (marked by λ̄2∗ ∗
ci = 0.002) with ω̄x,min =−0.8 at y* = 0.14
and z* = 0.98 (figure 16d). This vortex is spatially coincident with the detected Sta = 0.14.
The Str = 0.20 is associated with the hairpin vortices emanated from the separation bubble
near the leading edge of the roof, and Sta = 0.14 is linked to the oscillation of one pair of
longitudinal counter-rotating vortices behind the bubble, whose origins are the two foci at
913 A21-25
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
each end of the flow separation line (Zhang et al. 2015). When viewed in the y–z plane,
the vortex at y* = 0.14 is clockwise and the other at y* = −0.14 must be anticlockwise.
Therefore, the tangential velocity at the inner side of each longitudinal vortex about the
symmetry plane is upwards. Thus, at y* ∈ [−0.14, 0.14], the longitudinal vortices may act
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
to push up the hairpin vortices to the outer flow with a higher convection velocity to get
diffused, which may be responsible for the relatively weak peak of Str = 0.20 in Eu at F1
(y* = 0) (figure 16a).
The Eu measured over the sidewall at x* = −1.27 and y* = 0.63 (figure 16b) shows one
pronounced peak at f* = Sts = 0.27 over z* ∈ [0.14, 0.72] (G2 –G6 ). As above the roof, Sts
is associated with the side vortices emanated from the separation bubble near the leading
edge of the sidewall. At G4 (z* = 0.43) near the mid-height of the sidewall, this peak is less
pronounced, similarly to the case at F1 above the roof in the symmetry plane (figure 16a),
with the same flow physics involved, as is evident from one pair of counter-rotating
longitudinal vortices over the sidewall about the mid-height of the body (figure 16c,d).
Another predominant frequency of f* = Stl = 2.12 is detected at G1 –G2 near the lower edge
of the sidewall (figure 16b), which is ascribed to the vortices generated by the upstream
struts located at x* = −2.52 and y* = ±0.49 (Zhang et al. 2015).
In view of the strong three-dimensionality of the flow, the hot-wire measurements were
performed at various spanwise locations over the slanted surface, as shown in figure 17. A
peak at Str = 0.20 is discernible at H3 –H5 (x* = −0.36, y* ∈ [0.29, 0.43], z* = 0.91) near
the upper edge of the rear window and at I3 –I5 (x* = −0.09, y* ∈ [0.29, 0.43], z* = 0.91)
further downstream (figure 17a,b). However, this peak becomes less pronounced at H2 and
I2 (y* = 0.14) near the symmetry plane, and even vanishes at H1 and I1 (y* = 0), which
may be linked to the interaction between these structures with the longitudinal vortices
above the roof near the symmetry plane (figure 17a,c,d). In addition, Str is not detectable
in Eu measured at J1 –J4 and K1 –K5 at the height of the roof (z* = 0.86) (figure 17c,d) and
also at lower positions (not shown), indicating that these vortices travel above the upper
recirculation bubble.
A peak occurs at Sts = 0.27 in Eu measured at H6 , I6 , J6 and K6 near the side edge of
the rear window (y* = 0.57) (figure 17), where the C-pillar vortex takes place (figure 5).
A typical photograph captured in the x–y plane of z* = 0.72 over the rear window
(figure 18) exhibits nearly periodic structures right above the side edges of the rear window.
Obviously, this image should cut through the C-pillar vortices. Then, these nearly periodic
structures reconfirm Zhang et al.’s (2018) assertion that the side vortices may wrap around
and get tangled with the C-pillar vortex.
In order to detect possible nearly periodic structures formed in the gap between the
ground and the underside of the Ahmed body, hot-wire measurements were conducted
along L1 –L7 (x* = 0.09, z* = 0), downstream of the lower edge of the base (figure 19). Two
prominent frequencies occur at L3 –L7 (y* ∈ [0.36, 0.79]) in Eu . One is f* = Stg = 1.77 at
L3 –L5 over y* ∈ [0.36, 0.65], spatially corresponding to the gap vortex captured by Wang
et al. (2013). Another is f* = Stl = 2.12 at L6 –L7 over y* ∈ [0.72, 0.79], outside the lower
corner of the base, spatially coinciding with the lower vortex, which originates from the
vortices detected at G1 –G2 at the lower edge of the sidewall (figure 16b).
https://doi.org/10.1017/jfm.2020.1136
y∗ y∗
0.57 H6 0.57 I6
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
0.43 H5 0.43 I5
0.36 H4 0.36 I4
0.29 H3 0.29 I3
Str = 0.20 Str = 0.20
10–2 10–2
0.14 H2 0.14 I2
Eu 10–3 10–3
10–4 –2 0 H1 10–4 –2 0 I1
10 10–1 100 10 10–1 100
f∗ f∗
H6 J6 I1 I6
H1
K6
J1 K1
z
y
ϕ = 35° o x
y∗ y∗
0.57 J6 0.57 K6
Str = 0.20
0.43 J5
0.43 K5
0.36 J4 0.36 K4
0.29 J3
0.29 K3
occur when the ratio of the Hertz-averaged energy in Eu over St ∈ [Sti − 0.01; Sti + 0.01]
to that over St ∈ [Sti − 0.15; Sti + 0.15] exceeds 1.3, at which a pronounced peak can be
seen at Sti in Eu (Grandemange et al. 2013).
The hairpin vortices of Str = 0.20 from the roof are detectable at z* = 0.9 over y* ∈ [0.14,
0.36] in the plane of x* = 0.09 (figure 20a). Further downstream at x* = 0.20, 0.43 and
0.65 (figure 20b–d), these structures expand slightly downwards to z* = 0.81. The extent
913 A21-27
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
z∗ = 0.72
z
o
x
Figure 18. Typical photograph of the flow structure in the x–y plane of z* = 0.72 at Re = 0.22 × 105 . Flow is
from left to right.
where Str = 0.20 is detected shrinks in the region of 0.14 ≤ y* ≤ 0.22 and 0.81 ≤ z* ≤ 0.86
in the plane of x* = 0.86 (figure 20e), and vanishes completely beyond the plane of
x* = 1.29 (figure 20f,g), that is, the hairpin vortices disappear.
In the plane of x* = 0.09 (figure 20a), the side vortices of Sts = 0.27 are discernible
in two regions behind the side edge of the rear window and that of the base. However,
this frequency cannot be captured for z* ∈ [0.4, 0.6] about the sidewall, due to the
interaction of these vortices with the longitudinal vortices near the mid-height of the
sidewall (figure 16b–d). Beyond x* = 0.43 (figure 20c), the upper detections of Sts = 0.27
expand both downwards and inwards, covering most of the region above the upper edge
of the vertical base by x* = 0.65 (figure 20d); by x* = 1.29 (figure 20f ), the structures
spread over the central region and reach z* = 0.13 near the lower edge of the base. The
other region of Sts = 0.27 near the lower sidewall reaches the lower edge of the vertical
base at x* = 0.43 (figure 20c) and grows considerably around the lower corner of the base
from x* = 0.86 (figure 20e–g). As the vortices may interact with and wrap up around the
longitudinal trailing vortices, Sts = 0.27 is identifiable over x* ∈ [1.29, 1.55] in the wake
(figure 20f,g). This is quite different from the high-drag regime where only the unsteady
structures of St = 0.44 persist downstream of the recirculation bubbles over x* ∈ [1.4, 2.2]
(Zhang et al. 2015). In the high-drag regime, the vibrant C-pillar vortices interact with the
side vortices and tend to accelerate the breakup of the latter; while in the low-drag regime,
the weak ones would have a very small effect on the latter. Thus, the side vortices may
persist further downstream.
The structures of Stw = 0.30 emerge at x* = 0.43 around the mid-height and lower edge
https://doi.org/10.1017/jfm.2020.1136
of the base (figure 20c). The upper detections (z* > 0.20) are associated with the structures
from the upper recirculation bubble. The region enclosing the detections grows from
x* = 0.43 to 0.65 (figure 20c,d) but contracts appreciably beyond x* = 0.86 (figure 20e).
The structures vanish completely by x* = 1.29 (figure 20f ) where the unsteady structures
of Sts = 0.27 persist. The lower detections (z* < 0.20) of Stw = 0.30 associated with the
structures from the lower bubble expand transversely from x* = 0.43 to 0.86 (figure 20c–e)
913 A21-28
Flow structure around a low-drag Ahmed body
Stl = 2.12
y∗
0.79 L7
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
0.72 L6
0.65 L5
0.57 L4
Stg = 1.77
0.36 L3
10–2 0.29 L2
Eu
10–3
0 L1
10–4 –2
10 10–1 100
f∗
ϕ=
35°
y
o L7
L1
x
Figure 19. Power spectral density function Eu of the hot-wire signal measured along J1 –J7 (x* = 0.09,
z* = 0) in the wake, at Re = 0.62 × 105 .
and then spread upwards at x* = 0.86 (figure 20e). The region enclosing these detections
recedes, from x* = 0.86 to 1.55 (figure 20e–g), away from the centreline. While advected
downstream, the structures from the lower bubble may interact with the developing
longitudinal trailing vortices and be drawn transversely away from the plane of symmetry.
Apparently, the structures originated from the lower bubble persist longer than those from
the upper bubble, which is internally consistent with the observation from figure 13. Note
that Stw = 0.30 fails to be captured in the off-symmetry region of y* > 0.3 for x* < 1.29
(figure 20c–e). That is, the corresponding unsteady structures occur largely only around
the plane of y* = 0.
At x* = 0.09 (figure 20a), Stg = 1.77 associated with the gap vortices is detected around
the cylindrical support at y* ∈ [0.29, 0.65] and z* ∈ [−0.04, 0.04], while Stl = 2.12 is
captured outside the lower corner at y* = 0.72 and z* ∈ [0, 0.09]. The two frequencies
https://doi.org/10.1017/jfm.2020.1136
5. Reynolds-number effect
The information on St is crucially important for developing an effective active drag
reduction technique (e.g. Brunn & Nitsche 2006; Joseph et al. 2012; Barros et al. 2016b;
913 A21-29
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
0.6
o x
z∗ 0.5
x∗ 0 0.43 0.86 1.29 1.55 0.4
0.09 0.3
0.2
0.20 0.27 0.30 0.1
1.77 2.12 0
–0.1
0 0.2 0.4 0.6 0.8
(b) (c) (d )
0.9 0.9 0.9
0.8 0.8 0.8
0.7 0.7 0.7
0.6 0.6 0.6
0.5 0.5 0.5
z∗
0.4 0.4 0.4
0.3 0.3 0.3
0.2 0.2 0.2
0.1 0.1 0.1
0 0 0
–0.1 –0.1 –0.1
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
Tounsi et al. 2016). However, the relationship between Re and St in the low-drag regime
has not been well documented in the literature. In this section, the possible dependences
of the five predominant unsteady structures on Re are investigated based on hot-wire
measurements over Re ∈ [0.3, 2.7] × 105 . Figure 21 presents the variations of Str , Sts , Stw ,
913 A21-30
Flow structure around a low-drag Ahmed body
Str Sts Stg Stl Stw Present data
Stw Kohri et al. (2014)
2.4
Stl = 2.15
2.2
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
2.0
1.8
Stg = 1.84
St 1.6
Sts = 0.171 + 0.126 × Re/105
Stw = 0.30
0.4
0.2
Str = 0.143 + 0.070 × Re/105
Stg and Stl with Re, along with the data available in the literature for the same Ahmed body
of ϕ = 35° for comparison. A number of observations can be made.
Firstly, Stw varies little for the Re range examined, indicating a relatively weak Re
effect. Kohri et al. (2014) detected St ≈ 0.27 at Re = 0.73 × 105 in the symmetry plane
downstream of the lower edge of the base for the same Ahmed body (ϕ = 35°), which
is reasonably close to the present Stw (= 0.30). The slight deviation is probably due to
the differences in facilities, measurement techniques, etc., between the two experiments.
Note that Stw at ϕ = 25° increases nonlinearly from 0.44 to 0.54 from Re = 0.45 × 105 to
2.4 × 105 (Zhang et al. 2015), which is more appreciable than the present low-drag regime.
Vino et al. (2005) also observed in the high-drag regime (ϕ = 30°) that this frequency
increased from 0.36 to 0.39 over Re ∈ [4.40, 7.68] × 105 . The comparison suggests that
the Re effect on Stw is different between the low- and high-drag regimes. In their study
on the wake of an Ahmed body (ϕ = 25°), Joseph et al. (2012) observed a longitudinal
shrinkage in the separation region over the rear window as Re was increased from 4.5 × 105
to 9.0 × 105 . The shrunk separation region may influence significantly the downstream
reattaching flow over the rear window. Thacker et al. (2012) found from their PIV data
obtained in the wake of the Ahmed body (ϕ = 25°) that the flow separation over the
slanted surface was suppressed by rounding the upper edge of the rear window, which
led to a rise in the velocity magnitude of the reattaching flow before separating again at
the upper edge of the base. It is therefore inferred that, in the high-drag regime, the reduced
separation region over the slanted surface with increasing Re may increase the momentum
of flow separated from the upper edge of the base before being entrained into the upper
https://doi.org/10.1017/jfm.2020.1136
recirculation bubble, and hence affect Stw . In contrast, in the low-drag regime, the two
bubbles are little affected in size by Re due to their fixed separation points. As such, Stw
remains nearly unchanged at 0.3 in the Re range from 0.3 × 105 to 2.7 × 105 (figure 21).
The measured Str and Sts increase linearly for higher Re, fitting reasonably well
to Str = 0.143 + 0.070Re/105 and Sts = 0.171 + 0.126Re/105 , respectively. Zhang et al.
(2015) suggested that the frequency fj ( j = r or s) of the hairpin vortices is dictated
by fj ∼ U∞ /xRj (where xRj is the length of the bubble near the leading edge).
913 A21-31
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
Recirculation bubble Longitudinal vortices
Sta = 0.14
Hairpin vortices emanated
from the bubble
Str = 0.20
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
Weak C-pillar
vortex
Corner
vortex Upper
Side vortices recirculation
Sts = 0.27 bubble
Longitudinal
trailing vortex
Vortices generated
by the strut
Stl = 2.12 Lower vortex
Lower
recirculation
bubble Flow structure
Gap vortex Flow structure
Stg = 1.77 emanated from pinched off from
the upper bubble the lower bubble
Stw = 0.30
Ground floor
Figure 22. A conceptual model of the flow structure around the Ahmed model in the low-drag regime
(ϕ = 35°).
vortices and the upper recirculation bubble, along with the evolution and formation of the
trailing vortices. This model provides a full picture of the flow structures in the low-drag
regime, which covers steady and unsteady structures above the roof, over the sidewall and
the rear window, and behind the base, all based on the data in the literature as well as the
present data.
The flow reattaches soon after its separation near the leading edge of the roof,
forming a separation bubble that pulsates periodically (Minguez et al. 2008). One pair
913 A21-32
Flow structure around a low-drag Ahmed body
of predominantly longitudinal counter-rotating vortices takes place behind the bubble
(figure 16c,d), which is originated from two foci at the ends of the separation line (Spohn
& Gilliéron 2002). Owing to the bubble pulsation, the cores of these longitudinal vortices
oscillate at Sta ≈ 0.14 (figure 16a). Three-dimensional hairpin vortices are emanated from
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
the bubble and are advected downstream along the roof (Krajnović & Davidson 2005b;
Franck et al. 2009), which are responsible for Str = 0.20 (figure 16a). The longitudinal
vortex pair interacts with the hairpin vortices, accelerating the breakup of the latter near
the symmetry plane (figure 16a). By the same token, the side vortices of Sts = 0.27 are
emanated behind the separation bubble formed near the leading edge of each sidewall
(figure 16b). Another pair of predominantly longitudinal counter-rotating vortices are
generated from the two ends of this separation bubble (figure 16c,d). Apparently, the
flow over the side surface cannot be symmetric because of the presence of the ground.
Near the ground, flow rolls up about the lower edge of the sidewall under the pressure
difference between the flows inside and outside of the gap (Wang et al. 2013), forming
the lower vortex (figure 16c,d). This structure decays gradually downstream behind the
base (figure 6). Vortices are generated by the upstream struts and wrapped up by the lower
vortex, responsible for the detected Stl = 2.12 near the bottom of the sidewall (figure 16b).
The flow coming off the roof separates at the upper sharp edge of the rear window and
fails to reattach on the slanted surface, rolling up behind the base to form a big separation
bubble, covering both rear window and most of the upper part of the base (figure 2). The
C-pillar vortices are greatly weakened, as compared with the high-drag regime, due to a
reduced pressure difference between the flow over the rear window and that coming off
the sidewall. Their cores are deflected further away from the slanted surface as compared
with the high-drag case at ϕ = 25° (figure 5). The weak C-pillar vortices make the wake
much less three-dimensional than in the high-drag regime case (e.g. Ahmed et al. 1984;
Lienhart & Becker 2003).
Within the big upper recirculation bubble, an upwash flow near the upper base separates
from the corner between the rear window and the vertical base and rolls up (figures 2a
and 4b) under the pressure difference between the base flow and that over the rear window
(figure 4a) as well as the induction of the negative circulation within the bubble. Thus a
corner vortex is formed near the lower edge of the rear window, extending transversely
almost over the entire width of the body (figure 3). The upper recirculation bubble results
in a considerable rise in pressure on the slanted surface, compared with its counterpart in
the high-drag regime (Ahmed et al. 1984; Lienhart & Becker 2003), at a small expense
to bring down slightly the pressure on the upper base. On the other hand, there is a small
pressure drop near the lower end of the window where the corner vortex occurs (figure 4).
Zhang et al. (2018) deployed a combination of steady actuations along the upper and two
side edges of the rear window, and along the lower edge of the base to manipulate the flow
around an Ahmed body (ϕ = 25°). This manipulation acted to join the flow separation
bubble over the slanted surface and the upper recirculation bubble behind the base. The
resulting flow topology was quite similar to that in the low-drag regime (figure 2), though
without the presence of the corner vortex. As a result, drag reduction reached 26 % and the
corresponding drag coefficient was down to 0.267, even lower than its counterpart (0.32)
https://doi.org/10.1017/jfm.2020.1136
by 17 % in the low-drag regime at the same Re. It is therefore plausible that the corner
vortex may act as a partition to separate the high-pressure flow in the base region from the
relatively low-pressure flow over the entire slanted surface, playing a significant role in the
wake dynamics.
The nearly periodic structures pinched off from the lower recirculation bubble and those
emanated from the upper bubble are characterized by a phase shift of about 0.5π (figures 9,
10 and 13), not π as in the high-drag regime case (Zhang et al. 2015). These structures
913 A21-33
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
are characterized by a predominant frequency of Stw ≈ 0.30 for Re ∈ [0.3, 2.7] × 105
(figures 7, 20 and 21). The structures originated from the lower bubble persist longer than
those from the upper bubble (figures 13 and 20).
The upper recirculation bubble extends transversely about the symmetry plane, but tilts
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
downstream near each lateral side of the body under the effect of the streamwise flow
from the sidewall (figure 2), forming a horseshoe vortex (figures 3 and 6). The legs of this
horseshoe vortex develop inwards and downwards downstream. So do the pair of C-pillar
vortices (figure 6a–c). Further downstream, however, the relatively weak C-pillar vortices
merge into the horseshoe vortex legs and vanish quickly. It is the downstream tilting legs
of the upper recirculation bubble that form the trailing vortices (figure 6).
The structures formed above the roof and the side surface have a great impact on the
instability of the wake. Firstly, the hairpin vortices from the roof flow over the upper
recirculation bubble, which are responsible for the measured Str = 0.20 in the shear layer
separated from the upper edge of the rear window (figures 17 and 20). Secondly, the
side vortices may be wrapped up by the C-pillar vortices (figure 18), accounting for the
detected Sts = 0.27 above the side edge of the rear window (figure 17). This frequency
also shows up downstream of the base, even beyond x* = 1.55, coinciding with the trailing
vortices.
The gap vortices (Stg = 1.77) and the vortices from the upstream struts (Stl = 2.12) are
detected near the lower edge of the vertical base (figure 19). These vortices decay rather
quickly when convected downstream (figure 20).
Some remarks are due on the major differences between the low- and high-drag regimes.
Firstly, the two recirculation bubbles occur behind the base in the high-drag regime.
Flow separates from the upper edge of the rear window and then reattaches downstream,
forming a separation bubble over the rear window. It is flow separation from the upper and
lower edges of the base that forms the two recirculation bubbles. In the low-drag regime,
however, the formation of the upper recirculation bubble is totally different. A big upper
recirculation bubble is formed due to flow separating from the upper edge of the window
and shooting over the base. Within this bubble, the relatively high-pressure fluid behind
the base is recirculated towards the window, raising its pressure, contributing to the low
drag.
Secondly, the formation of the two trailing vortices is different. The C-pillar vortices in
the high-drag regime are very strong, snatching the lateral part of the upper recirculation
bubble behind the base and forming the trailing vortices (Ahmed et al. 1984; Venning
et al. 2017). In contrast, the C-pillar vortices are very weak in the low-drag regime and
vanish quickly. The trailing vortices originate from the downstream tilting legs of the upper
recirculation bubble (figure 6) and are much weaker in strength than their counterparts
in the high-drag regime. This weak strength implies a small magnitude of the dynamic
pressure, which is the source of the ‘vortex drag’ (Hucho & Sovran 1993), and hence
accounts, along with the relatively high pressure on the rear window due to the big
recirculation bubble, for the small drag in the low-drag regime.
Finally, there is a difference in the wake vortices; Stw remains nearly constant at
0.3 for Re ∈ [0.45, 2.4] × 105 in the low-drag regime but rises from 0.44 to 0.54 over
https://doi.org/10.1017/jfm.2020.1136
the same Re range in the high-drag regime (Zhang et al. 2015). The wake vortices are
emanated alternately from the two recirculation bubbles, which grow and burst. The upper
recirculation bubble in the low-drag regime originates from flow separation from a fixed
point, the upper edge of the rear window. The thus formed vortices are less Re-dependent.
In the high-drag regime, the separation region over the slanted surface contracts at the
higher Re, which is accompanied by an increase in the momentum of the downstream
913 A21-34
Flow structure around a low-drag Ahmed body
reattaching flow. The higher momentum of the flow separated from the upper edge of the
base accounts for a rise in Stw .
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
7. Conclusions
An experimental investigation has been conducted on the flow structure around a low-drag
Ahmed body with a slant angle of ϕ = 35° at Re ∈ [0.3, 2.7] × 105 . Extensive hot-wire
measurements have been conducted at more than 1500 points around the body. Flow
visualization, PIV and pressure measurements have also been performed. The data have
been carefully analysed. The following conclusions can be drawn out of this investigation.
A total of five distinct predominant frequencies have been detected in the wake, two
over the rear window and three behind the vertical base. Over the rear window, Str ∈
[0.17, 0.35] is detected in the region off the symmetry plane and higher than the roof.
The corresponding unsteady structures originate from the hairpin vortices emanated from
the separation bubble near the leading edge of the roof. Another unsteady structure of
Sts ∈ [0.21, 0.35] is detected near the sidewall of the model, referred to as the side vortex
emanated from the bubble near the leading edge of the sidewall, which wraps around
the weak C-pillar vortex. Behind the vertical base, the unsteady structures of Stg ≈ 1.77
and Stl ≈ 2.12 captured near the lower edge of the vertical base are attributed to vortices
generated by the downstream and upstream cylindrical struts, respectively. These four
unsteady structures are also captured in the high-drag regime (Zhang et al. 2015). Another
distinct predominant frequency of Stw ≈ 0.30 is captured in the central region behind the
vertical base, which is attributed to the nearly periodic structures emanated from the
upper recirculation bubble and those pinched off from the lower bubble. This structure
is different from that found in the high-drag regime.
Firstly, the nearly periodic structures are emanated at a significantly higher frequency,
Stw = 0.44 for Re = 0.62 × 105 , from two recirculation bubbles in the high-drag regime
(Zhang et al. 2015). Secondly, the emanated structures from the two bubbles are
characterized by a phase shift of about 0.5π in the low-drag regime, while those in the
high-drag regime are in antiphase (Zhang et al. 2015). This difference is connected to
the marked disparity in the size and strength of the upper recirculation bubble between
the two regimes. This non-antiphase behaviour between the structures emanated from the
two bubbles accounts for the short lifespan of these structures in the low-drag regime.
Finally, Stw increases with increasing ϕ in the high-drag regime but remains constant in
the low-drag regime. It is found that Stw scales with the vertical distance S between the
separation points of the two recirculation bubbles. As the effective separation angle θ of
the separated shear layer that forms the upper bubble may affect Stw , the characteristic
length may be corrected by αS (α = (90° − θ)/90°), which reflects physically the bubble
size. A small S or a large θ corresponds to a small bubble size and subsequently a large
Stw . As a result, all Stw data collapse to Stw+ ≈ 0.20, irrespective of the value of ϕ or the
flow regimes.
A small corner vortex is formed near the lower edge of the rear window. The pressure
https://doi.org/10.1017/jfm.2020.1136
on the base is higher than that on the rear window. This pressure disparity, along with the
induction of the negative circulation of the big bubble, drives an upwash and backward
flow, forming this corner vortex. This structure plays a significant role in the wake
dynamics, acting to separate the high-pressure flow behind the base from the low-pressure
flow over the rear window.
The Reynolds-number effects on these unsteady structures have been investigated. The
effect of Re on Stw ≈ 0.30 is negligibly small over the Re range examined (Re ∈ [0.3,
913 A21-35
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
2.7] × 105 ). The dependences of Str , Sts , Stl and Stg on Re are similar to their counterparts
in the high-drag regime because the origins of the corresponding unsteady structures are
the same for the two regimes.
Based on the present data and those reported in the literature, a conceptual model is
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
proposed for the flow structure around an Ahmed body in the low-drag regime, which
provides a rather comprehensive picture for the complex flow. The present findings
are incorporated into this model, including the wake vortices of Stw ≈ 0.30 and their
formation process, the corner vortex and its dynamical role, the interaction between the
C-pillar vortex and the upper recirculation bubble leg, and the formation and development
of the longitudinal trailing vortex. Previously reported coherent structures, such as the
hairpin vortices, side vortices, gap vortices and lower vortices, are also included. The
dimensionless frequencies are identified for the unsteady structures.
Funding. Y.Z. wishes to acknowledge support given to him from NSFC through grants 11632006
and 91952204 and from the Research Grants Council of Shenzhen Government through grant
JCYJ20190806143611025. B.F.Z. is grateful for the financial support from NSFC through grant 11902102.
Z.Y.C. acknowledges the financial support from NSFC through grant 11772140.
Author ORCIDs.
K. Liu http://orcid.org/0000-0003-1362-961X;
B.F. Zhang http://orcid.org/0000-0001-8696-5169.
R EFERENCES
A HMED , S.R., R AMM , G. & FALTIN , G. 1984 Some salient features of the time-averaged ground vehicle
wake. SAE Technical Paper 840300, pp. 1–30. Society of Automotive Engineers, Inc., Warrendale, PA.
A NTONIA , R.A., Z HOU, Y. & M ATSUMURA , M. 1993 Spectral characteristics of momentum and heat-transfer
in the turbulent wake of a circular-cylinder. Exp. Therm. Fluid Sci. 6, 371–375.
BALACHANDAR , S., M ITTAL , R. & NAJJAR , F.M. 1997 Properties of the mean recirculation region in the
wakes of two-dimensional bluff bodies. J. Fluid Mech. 351, 167–199.
BARROS , D., BORÉE , J., C ADOT, O., S POHN , A. & N OACK , B.R. 2017 Forcing symmetry exchanges and
flow reversals in turbulent wakes. J. Fluid Mech. 829, R1.
BARROS , D., BORÉE , J., N OACK , B.R. & S POHN , A. 2016a Resonances in the forced turbulent wake past a
3D blunt body. Phys. Fluids 28, 065104.
BARROS , D., BORÉE , J., N OACK , B.R., S POHN , A. & RUIZ , T. 2016b Bluff body drag manipulation using
pulsed jets and Coanda effect. J. Fluid Mech. 805, 422–459.
B EN C HIEKH , M., M ICHARD , M., G ROSJEAN , N. & B ERA , J. 2004 Reconstruction temporelle d’un champ
aérodynamique instationnaire à partir de mesures PIV non résolues dans le temps. In Proceedings of 9ème
Congrès Francophone de Vélocimétrie Laser, Brussels, Belgium, September 2004, paper D.8.
B RUNEAU, C.-H., C REUSÉ , E., D EPEYRAS , D., G ILLIÉRON , P. & M ORTAZAVI , I. 2011 Active procedures
to control the flow past the Ahmed body with a 25° rear window. Intl J. Aerodyn. 1, 299–317.
B RUNN , A. & N ITSCHE , W. 2006 Active control of turbulent separated flows over slanted surfaces. Intl J.
Heat Fluid Flow 27, 748–755.
C HOI , H., L EE , J. & PARK , H. 2014 Aerodynamics of heavy vehicles. Annu. Rev. Fluid Mech. 46, 441–468.
C HONG , M.S., P ERRY, A.E. & C ANTWELL , B.J. 1990 A general classification of three-dimensional flow
fields. Phys. Fluids A: Fluid Dyn. 2, 765–777.
https://doi.org/10.1017/jfm.2020.1136
C ONAN , B., A NTHOINE , J. & P LANQUART, P. 2011 Experimental aerodynamic study of a car-type bluff body.
Exp. Fluids 50, 1273–1284.
FARES , E. 2006 Unsteady flow simulation of the Ahmed reference body using a lattice Boltzmann approach.
Comput. Fluids 35, 940–950.
F RANCK , G., N IGRO , N., S TORTI , M. & D’ ELIA , J. 2009 Numerical simulation of the flow around the
Ahmed vehicle model. Latin Am. Appl. Res. 39, 295–306.
G RANDEMANGE , M., G OHLKE , M. & C ADOT, O. 2013 Turbulent wake past a three-dimensional blunt body.
Part 1. Global modes and bi-stability. J. Fluid Mech. 722, 51–84.
913 A21-36
Flow structure around a low-drag Ahmed body
G UILMINEAU, E. 2008 Computational study of flow around a simplified car body. J. Wind Engng Ind. Aerodyn.
96, 1207–1217.
G UILMINEAU, E., D ENG , G.B., L EROYER , A., Q UEUTEY, P., V ISONNEAU, M. & WACKERS , J. 2017
Assessment of hybrid RANS-LES formulations for flow simulation around the Ahmed body. Comput.
Fluids 176, 302–319.
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
H UCHO , W.H. & S OVRAN , G. 1993 Aerodynamics of road vehicles. Annu. Rev. Fluid Mech. 25, 485–537.
H USSAIN , A.K.M.F. & R EYNOLDS , W.C. 1970 The mechanics of an organized wave in turbulent shear flow.
J. Fluid Mech. 41, 241–258.
J OSEPH , P., A MANDOLESE , X. & A IDER , J.L. 2012 Drag reduction on the 25 degrees slant angle Ahmed
reference body using pulsed jets. Exp. Fluids 52, 1169–1185.
KOHRI , I., YAMANASHI , T., NASU, T., H ASHIZUME , Y. & K ATOH , D. 2014 Study on the transient behaviour
of the vortex structure behind Ahmed body. SAE Intl J. Passeng. Cars - Mech. Syst. 7, 586–602.
K RAJNOVI Ć , S. & DAVIDSON , L. 2005a Flow around, a simplified car, part 1: large eddy simulation. Trans.
ASME J. Fluids Engng 127, 907–918.
K RAJNOVI Ć , S. & DAVIDSON , L. 2005b Flow around a simplified car, part 2: understanding the flow. Trans.
ASME J. Fluids Engng 127, 919–928.
L AHAYE , A., L EROY, A. & KOURTA , A. 2014 Aerodynamic characterisation of a square back bluff body flow.
Intl J. Aerodyn. 4, 43–60.
L EHUGEUR , B., G ILLIÉRON , P. & TA-P HUOC , L. 2005 Characterization of longitudinal vortices in the wake
of a simplified car model. In 23rd AIAA Applied Aerodynamics Conference, AIAA Paper 2015-5383.
L IENHART, H. & B ECKER , S. 2003 Flow and turbulence structure in the wake of a simplified car model. SAE
Technical Paper 2003-01-0656, Society of Automotive Engineers, Inc, Warrendale, PA.
L IENHART, H., S TOOTS , C. & B ECKER , S. 2002 Flow and turbulence structures in the wake of a simplified car
model (Ahmed model). In New Results in Numerical and Experimental Fluid Mechanics III, pp. 323–330.
Springer.
M EILE , W., L ADINEK , T., B RENN , G., R EPPENHAGEN , A. & F UCHS , A. 2016 Non-symmetric bi-stable
flow around the Ahmed body. Intl J. Heat Fluid Flow 57, 34–47.
M INGUEZ , M., PASQUETTI , R. & S ERRE , E. 2008 High-order large-eddy simulation of flow over the “Ahmed
body” car model. Phys. Fluids 20, 095101.
Ö STH , J., N OACK , B.R., K RAJNOVI Ć , S., BARROS , D. & BORÉE , J. 2014 On the need for a nonlinear
subscale turbulence term in POD models as exemplified for a high-Reynolds-number flow over an Ahmed
body. J. Fluid Mech. 747, 518–544.
VAN O UDHEUSDEN , B.W., S CARANO , F., VAN H INSBERG , N.P. & WATT, D.W. 2005 Phase-resolved
characterization of vortex shedding in the near wake of a square-section cylinder at incidence. Exp. Fluids
39, 86–98.
R AO , A., M INELLI , G., BASARA , B. & K RAJNOVI Ć , S. 2018 On the two flow states in the wake of a
hatchback Ahmed body. J. Wind Engng Ind. Aerodyn. 173, 262–278.
ROUMÉAS , M., G ILLIÉRON , P. & KOURTA , A. 2009 Drag reduction by flow separation control on a car after
body. Intl J. Numer. Meth. Fluids 60, 1222–1240.
S CHLICHTING , H. & G ERSTEN , K. 2000 Boundary Layer Theory, 8th edn, p. 22. Springer.
S CIACCHITANO , A., W IENEKE , B. & S CARANO , F. 2013 PIV uncertainty quantification by image matching.
Meas. Sci. Technol. 24, 045302.
S ELLAPPAN , P., M CNALLY, J. & A LVI , F.S. 2018 Time-averaged three-dimensional flow topology in the
wake of a simplified car model using volumetric PIV. Exp. Fluids 59, 124.
S IMS -W ILLIAMS , D. 2001 Self-excited aerodynamic unsteadiness associated with passenger cars. Doctoral
dissertation, Durham University.
S IROVICH , L. 1987 Turbulence and the dynamics of coherent structures. I. Coherent structures. Q. Appl. Maths
45, 561–571.
S POHN , A. & G ILLIÉRON , P. 2002 Flow separations generated by a simplified geometry of an automotive
vehicle. In IUTAM Symposium on Unsteady Separated Flows. Kluwer Academic.
S TRACHAN , R.K., K NOWLES , K. & L AWSON , N.J. 2007 The vortex structure behind an Ahmed reference
model in the presence of a moving ground plane. Exp. Fluids 42, 659–669.
https://doi.org/10.1017/jfm.2020.1136
T HACKER , A., AUBRUN , S., L EROY, A. & D EVINANT, P. 2012 Effects of suppressing the 3D separation on
the rear slant on the flow structures around an Ahmed body. J. Wind Engng Ind. Aerodyn. 107, 237–243.
T HACKER , A., AUBRUN , S., L EROY, A. & D EVINANT, P. 2013 Experimental characterization of flow
unsteadiness in the centerline plane of an Ahmed body rear slant. Exp. Fluids 54, 1479.
T OUNSI , N., M ESTIRI , R., K EIRSBULCK , L., O UALLI , H., H ANCHI , S. & A LOUI , F. 2016 Experimental
study of flow control on bluff body using piezoelectric actuators. J. Appl. Fluid Mech. 9, 827–838.
913 A21-37
K. Liu, B.F. Zhang, Y.C. Zhang and Y. Zhou
T UNAY, T., SAHIN , B. & O ZBOLAT, V. 2014 Effects of rear slant angles on the flow characteristics of Ahmed
body. Exp. Therm. Fluid Sci. 57, 165–176.
T UNAY, T., YANIKTEPE , B. & SAHIN , B. 2016 Computational and experimental investigations of the vortical
flow structures in the near wake region downstream of the Ahmed vehicle model. J. Wind Engng Ind.
Aerodyn. 159, 48–64.
Downloaded from https://www.cambridge.org/core. Robert Gordon University, on 28 May 2021 at 20:32:10, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms.
V ENNING , J., L O JACONO , D., B URTON , D., T HOMPSON , M.C. & S HERIDAN , J. 2017 The nature of the
vortical structures in the near wake of the Ahmed body. Proc. Inst. Mech. Engrs D: J. Automob. Engng
231, 1239–1244.
V INO , G., WATKINS , S., M OUSLEY, P., WATMUFF , J. & P RASAD , S. 2005 Flow structures in the near-wake
of the Ahmed model. J. Fluids Struct. 20, 673–695.
VOLPE , R., D EVINANT, P. & KOURTA , A. 2015 Experimental characterization of the unsteady natural wake
of the full-scale square back Ahmed body: flow bi-stability and spectral analysis. Exp. Fluids 56, 99.
WANG , X.W., Z HOU, Y., P IN , Y.F. & C HAN , T.L. 2013 Turbulent near wake of an Ahmed vehicle model.
Exp. Fluids 54, 1490.
W EN , X., TANG , H. & D UAN , F. 2015 Vortex dynamics of in-line twin synthetic jets in a laminar boundary
layer. Phys. Fluids 27, 083601.
Z HANG , B.F., L IU, K., Z HOU, Y., T O , S. & T U, J.Y. 2018 Active drag reduction of a high-drag Ahmed body
based on steady blowing. J. Fluid Mech. 856, 351–396.
Z HANG , B.F., Z HOU, Y. & T O , S. 2015 Unsteady flow structures around a high-drag Ahmed body. J. Fluid
Mech. 777, 291–326.
Z HOU, J., A DRIAN , R.J., BALACHANDAR , S. & K ENDALL , T.M. 1999 Mechanisms for generating coherent
packets of hairpin vortices in channel flow. J. Fluid Mech. 387, 353–396.
Z HOU, Y., D U, C., M I , J. & WANG , X.W. 2012 Turbulent round jet control using two steady minijets. AIAA
J. 50, 736–740.
https://doi.org/10.1017/jfm.2020.1136
913 A21-38