You are on page 1of 46

Progress in Aerospace Sciences xxx (2017) 1–46

Contents lists available at ScienceDirect

Progress in Aerospace Sciences


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

A review of recent developments in the understanding of transonic


shock buffet
Nicholas F. Giannelis a, *, Gareth A. Vio a, Oleg Levinski b
a
School of Aerospace, Mechanical and Mechatronic Engineering, The University of Sydney, NSW 2006, Sydney, Australia
b
Aerodynamics and Aeroelasticity, Aerospace Division, Defence Science and Technology Group, Melbourne, VIC 3207, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: Within a narrow band of flight conditions in the transonic regime, interactions between shock-waves and
Transonic shock buffet intermittently separated shear layers result in large amplitude, self-sustained shock oscillations. This phenome-
Unsteady aerodynamics non, known as transonic shock buffet, limits the flight envelope and is detrimental to both platform handling
Shock-wave/boundary layer interaction
quality and structural integrity. The severity of this instability has incited a plethora of research to ascertain an
Nonlinear aeroelasticity
underlying physical mechanism, and yet, with over six decades of investigation, aspects of this complex phe-
Shock control
nomenon remain inexplicable. To promote continual progress in the understanding of transonic shock buffet, this
review presents a consolidation of recent investigations in the field. The paper begins with a conspectus of the
seminal literature on shock-induced separation and modes of shock oscillation. The currently prevailing theories
for the governing physics of transonic shock buffet are then detailed. This is followed by an overview of
computational studies exploring the phenomenon, where the results of simulation are shown to be highly sen-
sitive to the specific numerical methods employed. Wind tunnel investigations on two-dimensional aerofoils at
shock buffet conditions are then outlined and the importance of these experiments for the development of
physical models stressed. Research considering dynamic structural interactions in the presence of shock buffet is
also highlighted, with a particular emphasis on the emergence of a frequency synchronisation phenomenon. An
overview of three-dimensional buffet is provided next, where investigations suggest the governing mechanism
may differ significantly from that of two-dimensional sections. Subsequently, a number of buffet suppression
technologies are described and their efficacy in mitigating shock oscillations is assessed. To conclude, recom-
mendations for the direction of future research efforts are given.

1. Introduction suction surfaces of an aerofoil. Through the investigations of Mabey [2]


and Gibb [3], a working model of Type I buffet was developed, whereby
Within a narrow region of the transonic flight regime, the interactions shock-wave/boundary layer interactions on both surfaces initiate phase-
between shock-waves and thin, separated shear layers give rise to large locked shock oscillations in opposing directions. As the shock on the
amplitude, autonomous shock oscillations. This instability, commonly upper surface moves upstream, it weakens. This permits reattachment of
known as transonic shock buffet, acts as a limiting factor in aircraft the separated zone and propels the shock downstream. The shock motion
performance. The reduced frequency of shock oscillation is typically on on the lower surfaces occurs in an identical manner, with a 180 phase
the order of the low-frequency structural modes, resulting in an aircraft shift, yielding self-sustained shock buffet cycle. As Type I buffet is criti-
that is susceptible to limit cycle oscillations (LCOs), and as a conse- cally dependent on the shock having sufficient strength to produce sep-
quence, diminished handling quality and fatigue life. aration, several authors have proposed the prediction of buffet onset by
Hilton & Fowler [1] first observed transonic shock-induced oscilla- the Mach number immediately ahead of the shock [2–4].
tions over six decades ago, yet the physics governing aspects of this Type II shock buffet is characteristic of modern supercritical aerofoils
complex phenomenon remains elusive. Various numerical and experi- and involves upper surface shock oscillations at non-zero angles of attack.
mental investigations have identified two distinct types of shock buffet A working model of this second type that is unequivocally accepted by
on aerofoils. Type I buffet typically occurs at zero incidence on biconvex the research community has yet to be determined. Early work by Pearcey
sections and encompasses shock oscillations on both the pressure and [5,6], Pearcey & Holder [7] and Pearcey et al. [8] was instrumental in

* Corresponding author.
E-mail address: nicholas.giannelis@sydney.edu.au (N.F. Giannelis).

http://dx.doi.org/10.1016/j.paerosci.2017.05.004
Received 1 March 2017; Received in revised form 22 May 2017; Accepted 25 May 2017
Available online xxxx
0376-0421/© 2017 Elsevier Ltd. All rights reserved.

Please cite this article in press as: N.F. Giannelis, et al., A review of recent developments in the understanding of transonic shock buffet, Progress in
Aerospace Sciences (2017), http://dx.doi.org/10.1016/j.paerosci.2017.05.004
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Nomenclature ~
d modified length scale
fd delaying function
M freestream Mach number xþ nondimensional streamwise distance
α freestream angle of attack/vortex generator pitch angle zþ nondimensional spanwise distance
Re chord-based Reynolds number τp period downstream propagation
xs mean shock location τu period upstream propagation
ap downstream pressure perturbation convection velocity τu;u period upstream propagation above upper surface
au upstream pressure perturbation convection velocity τu;l period upstream propagation below lower surface
c chord fsb shock buffet frequency
b span fα pitch natural frequency
τ buffet period/time delay fh heave natural frequency
ϕ phase fα0 wind-off pitch natural frequency
a speed of sound fh0 wind-off heave natural frequency
Mc local Mach number ω reduced frequency
R constant ζ structural damping
Ms upper surface Mach number V reduced velocity
ρ freestream density Λ sweep angle
q two-dimensional flow state vector fρ; u; v; T; ~νg δ TED deflection
u streamwise velocity δ mean TED deflection
v transverse velocity A TED amplitude
T temperature P pressure
t time β trailing edge flap deflection/vortex generator skew angle
ω frequency Cl0 balanced lift coefficient
S blending function λ dimensionless controller gain/spanwise vortex
CL lift coefficient generator spacing
f frequency h vortex generator height
yþ nondimensional wall-normal distance l vortex generator length
~ν eddy viscosity d fluidic vortex generator orifice diameter
d wall distance (length scale) Cμ momentum coefficient
~
S local deformation rate θramp shock control bump ramp angle
Δ grid size (¼ maxðΔx ; Δy ; Δz Þ) ltail shock control bump tail length
CDES constant

characterising the various forms of upper surface separation, particularly downstream excursion. Type C motion is qualitatively distinct from the
shock-induced separation bubbles, experienced by conventional aerofoils preceding modes. The shock travels upstream, initially strengthening and
at transonic conditions. Two distinct modes of separation were identified; then weakening, but continuing to move forward, eventually propagating
Model A consisting only of a shock-induced separation bubble and Model forward into the oncoming flow as a free shock-wave. Although these
B for which trailing edge separation is either additionally present or shock motions were originally identified with oscillating aerofoils, each
incipient. Three variants of Model B were also identified; rear separation has subsequently been observed in rigid wing sections at certain flight
provoked by the formation of a bubble, rear separation provoked by the conditions [12].
shock and a third in which rear separation is present from the outset. Considering Tijdeman Type A [11] shock motions, Lee [13] proposed
The investigations by Pearcey and his co-authors culminated in the an acoustic wave-propagation feedback model as the underlying mech-
first model for the prediction of buffet onset in Type II shock oscillations; anism governing the autonomous shock oscillations. In this model, the
a relationship between trailing edge pressure divergence and large-scale motion of the shock-wave generates downstream propagating pressure
unsteadiness. For aerofoils in which separation bubbles are present, waves, with the instability growing as it travels from the separation point
Pearcey [6] and Pearcey & Holder [7] related the onset of buffet to the through the shear layer. The separated flow induces a de-cambering ef-
Mach number or angle of attack for which the separation bubble extends fect, and interactions with the flow at the trailing edge produce pressure
to the trailing edge and bursts. This bubble bursting mechanism governing waves that travel upstream in the subsonic flow above the boundary
buffet onset is easily identified through the divergence of trailing edge layer. Interaction between these upstream propagating disturbances and
pressure. Although bubble bursting as the cause of onset was initially the shock completes a feedback loop, yielding sustained shock motion.
supported by experimental and computational findings, recent in- Analogous to the bubble bursting mechanism of Pearcey [6], conflicting
vestigations have produced conflicting evidence [9,10] and bubble evidence has been presented in literature regarding the validity of Lee's
bursting is now widely discounted as a potential mechanism governing [13] model.
shock buffet. A mechanism underlying Tijdeman Type B [11] shock oscillations on
In the seminal work of Tijdeman [11], three distinct modes of shock the NACA 0012 aerofoil based on an unstable shock-wave/separation
motion were characterised experimentally by observing the effects of bubble interaction has also been proposed by Raghunathan et al. [14].
sinusoidal flap deflections on the NACA 64A006 aerofoil. Type A shock The authors highlight that the shock strength must be sufficient to induce
motion is represented by near sinusoidal shock oscillations across the a separation bubble. The appearance of this separation bubble initiates
upper surface of the aerofoil, for which the shock is present throughout periodic motion of the shock, which is sustained through the alternating
the entire buffet cycle but varies in strength, with maximum shock expansion and collapse of the bubble on the upper aerofoil surface.
strength achieved during the upstream excursion. Type B motion re- Throughout the cycle, the varying extent of the separated region acts to
sembles Type A; however, the magnitude of shock strength variation is change the effective camber of the aerofoil, with the trailing edge playing
considerably larger, resulting in a disappearance of the shock during the an integral role in communicating flow states between the suction and

2
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

pressure surfaces. from the shock foot. The motion of the shock-wave generates down-
The understanding of transonic shock buffet described thus far re- stream propagating pressure waves of velocity ap , with the instability
flects a classical perspective, encompassing much of the early experi- growing as it travels from the separation point through the separated
mental and numerical investigations that sought to identify the shear layer. As the disturbances reach the trailing edge, upstream prop-
underlying flow physics. A comprehensive review of this early work is agating pressure waves are produced such that the unsteady Kutta con-
provided by Lee [12]. The purpose of the present paper is to provide an dition is satisfied. These Kutta waves travel towards the shock in the
overview of recent developments in the understanding of transonic shock subsonic region above the separated flow at a velocity au . The interaction
buffet; research that has followed the review of Lee [12]. Where neces- between the upstream propagating pressure waves and the shock results
sary for context, an overview of the classical work will be presented, in an energy exchange, completing the feedback loop and sustaining the
however, it is assumed the reader is broadly familiar with the various periodic shock oscillation.
aspects of transonic shock oscillations outlined by Lee [12]. Additionally, The appeal of Lee's [13] model stems widely from the ability to pre-
this review is limited to Type II shock oscillations on aerofoils and wings. dict the shock oscillation frequency with a simple relationship directly
The justification of this limited scope is twofold. Firstly, as discussed, a related to observable variables. As the shock cycle is dependent on the
working model of Type I shock buffet has been developed by Mabey [2] time taken for disturbances to propagate downstream to the trailing edge
and Gibb [3]. Secondly, the majority of recent literature pertains to issues and then again upstream to the shock, the complete shock period is
encountered by civil transport aircraft, namely, aircraft with supercritical proposed as the sum of these two propagation times:
aerofoil sections that are susceptible to upper surface shock oscillations.
c  x
The remainder of this paper is organised as follows: Section 2 pro- τ ¼ ∫ xs 1 ap dx  ∫ cs 1=au dx (1)
vides an overview of research considering the governing physics under-
lying the transonic shock buffet phenomenon, including a description of where τ is the period of the buffet cycle, xs is the mean shock location and
Lee's [13] wave-propagation feedback model. In Section 3, numerical c is the chord. To validate the model, data from transonic wind tunnel
studies of transonic shock oscillations are reviewed, with a particular experiments conducted by Lee [15] on the BGK No. 1 supercritical
emphasis on the sensitivity of simulations to turbulence modelling, aerofoil at Re ¼ 20  106 are employed. In Fig. 2, the magnitude and
spatial and temporal discretisation and numerical schemes in Unsteady phase diagrams of the pressure signals (with respect to the shock motion)
Reynolds-Averaged Navier-Stokes (URANS) computations and the from this experiment at M ¼ 0:746 and α ¼ 6:066 are shown. The
applicability of scale-resolving methods. Recent experimental in- contributions to magnitude and phase from the fundamental and first
vestigations of transonic shock buffet on aerofoils are then discussed in harmonic frequency are decomposed, and as the magnitude of the first
Section 4. In Section 5, an area which has received particular attention harmonic is comparatively small, the model is developed based on the
over the past decade, dynamic interactions in the presence of shock behaviour of the fundamental frequency. Evident in Fig. 2(b), the phase
buffet, is examined. Explicit focus is given to the relationship between angle of the fundamental frequency varies approximately linearly behind
shock buffet as an aerodynamic resonance phenomenon and the large the shock. Nonetheless, this is not representative of all conditions
amplitude structural oscillations that follow from this resonance. Prog- considered, with the slope of the phase dϕ=dx not typically constant for
ress towards understanding the physics governing three-dimensional the BGK No. 1 aerofoil. It is this phase relationship that is applied to
shock buffet is outlined in Section 6, followed by a description of determine the velocity ap of the downstream propagating pressure waves.
buffet suppression technologies in Section 7. Some concluding remarks The upstream propagation velocity au is computed by:
and the author's perspective on critical aspects of the transonic buffet
phenomenon that have yet to be addressed in open literature are then au ¼ ð1  Mc Þa (2)
provided in Section 8.
where a is the local speed of sound and Mc is the local Mach number of
2. Governing physics the flow behind the shock, computed in accordance with Tijdeman
[11] by:
2.1. Wave-propagation feedback
Mc ≈RðMs  MÞ þ M (3)
Lee [13] proposed a model that enabled the prediction of shock
where M and Ms are the freestream and upper surface aerofoil Mach
oscillation frequency for Tijdeman [11] Type A instabilities. In Lee's [13]
numbers, respectively and R acts as a relaxation factor (0.7 to achieve
model, the periodic shock motions are a consequence of an acoustic
good correlations with the BGK No. 1 experiments). It is important to
wave-propagation feedback mechanism, which is shown graphically in
note that calculation of ap and au is not limited to the approach described
Fig. 1 for a symmetric aerofoil with shock-induced separation emanating
by Lee, with alternative methods provided by Erickson & Stephenson
[16], Mabey [2] and Mabey et al. [4].
The predictions made by Lee's [13] model for the shock oscillation
frequencies of the BGK No. 1 aerofoil are in fair agreement with the
values computed experimentally through force balance spectra, partic-
ularly considering the uncertainties related to shock location and Mc . In
his original work, Lee [13] found the model yields the most accurate
predictions at higher Mach numbers and incidence.
Although Lee's [13] model demonstrated fair agreement to experi-
ments of the BGK No. 1 aerofoil, subsequent literature has been some-
what conflicting regarding the applicability of the original formulation of
the wave-propagation feedback model. In a URANS analysis of the BGK
No. 1 aerofoil at shock buffet conditions, Xiao et al. [17] reported
excellent agreement between reduced shock frequencies computed
through fast Fourier transform of the lift signal and that provided by Lee's
[13] model. The improved predictions relative to Lee's [13] original work
may be attributed to the direct availability of the wave speeds, circum-
Fig. 1. Model of self-sustained shock oscillation (adapted from Lee [13]). venting the need for empirical correlations. With the entire unsteady

3
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

2.2. Global aerodynamic mode instability

A competing theory for the mechanism governing transonic shock


buffet, related to the instability in a global aerodynamic mode, has been
posited by Crouch et al. [10,21,22]. Global mode decomposition of the
steady transonic flowfield at pre-buffet conditions reveals a marginally
stable eigenvalue related to the streamwise velocity component. With
increasing incidence, a Hopf bifurcation results in this eigenvalue
crossing the stability boundary, yielding phase locked motion of the
shock and separated boundary layer. In this section, the method
employed by Crouch et al. [21] is detailed.

2.2.1. Formulation of the linearised system


The method posed by Crouch et al. [21] considers the two-
dimensional, viscous, compressible RANS equations; encompassing
continuity, streamwise momentum, transverse momentum, energy and
eddy viscosity, expressed in terms of the state vector:

q ¼ fρ; u; v; T; ~νg (4)

where ρ represents density, u and v are the streamwise and transverse


velocities respectively, T is temperature and ~ν the modified viscosity. The
state vector q is then decomposed into mean (q) and fluctuating (q0 )
components, such that:

q ¼ q þ q0 (5)
For conditions sufficiently close to a steady state, linearisation of the
governing RANS equations proceeds by assuming the fluctuating
component q0 behaves as a small perturbation to the mean flow q. The
complete set of linearised equations are omitted here for brevity, how-
ever they can be expressed in the simplified operator form:


A½q0  þ Bq ½q0  ¼ 0 (6)
∂t

where A is a linear operator comprised of the time derivative terms of the


RANS equations and Bq is a linear operator including the linear terms of
the RANS equations and the terms resulting from the nonlinear coupling
between q and q0 .
The perturbations to the mean flowfield q may then be described by
time-harmonic aerodynamic modes:

q0 ðx; y; tÞ ¼ b
q ðx; yÞ⋅eiωt (7)

where bq is an eigenfunction reflecting the mode shape and ω is the fre-


Fig. 2. Magnitude and phase of pressure waves propagating downstream in separated
flow region (Lee [115]).
quency. Substitution of Equation (7) into the system of Equation (6) and
multiplication by a conditioning matrix produces the final system
of equations:

flowfield known, the wave speeds and direction are computed directly iωb
q þ LðqÞ⋅b
q¼0 (8)
through two-point cross-correlations of pressure fluctuations on the
aerofoil surface and within the separated flow region. Similarly, Deck where L is a second-order differential operator. Similarly, Equation (5) is
[18] saw excellent agreement in a Zonal Detached-Eddy Simulation of passed through the appropriate RANS boundary conditions and Riemann
transonic flow over the OAT15A aerofoil, with differences in the invariants, resulting in an eigenvalue problem to be solved for the
computed buffet frequencies on the order of 5%. Nonetheless, in a sub- complex frequency ω and aerodynamic mode shape b q.
sequent study by Garnier & Deck [19], LES simulation of flow around the
OAT15A indicated discrepancies between Lee's [13] model and experi- 2.2.2. Method of solution
ment on the order of 60%. Jacquin et al. [20] came to analogous findings, Crouch et al. [21] employed a finite difference approximation for the
with Lee's [13] model performing poorly at computing the buffet fre- solution of Equation (8) and the associated boundary conditions and
quency of the OAT15A aerofoil. A modified model was suggested by the invariants. Discretisation of the steady RANS equations is performed with
authors (detailed in Section 4), whereby the upstream propagating Roe's scheme [23], using the third-order κ scheme [24] for the inviscid
pressure waves also travel along the lower surface and around the leading fluxes and second-order central differencing for the viscous and thermal
edge, impinging the shock from both upstream and downstream di- fluxes. A first-order upwind scheme is used for the turbulent convective
rections. Although the modified model improved the frequency pre- quantities. Interpolation of the state vector and cell coefficient matrix is
dictions, a difference of 36% remained, an indication that the model of performed through a blended third-order upwind/fourth-order central
Lee [13] is not robust to changes in aerofoil geometry. difference scheme, to reduce the numerical dissipation induced by the
upwind differencing.

4
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

The authors give particular attention to the treatment of shocks. Due


to the linearisation of the RANS equations, small unsteady perturbations
to the flowfield may yield fictitiously large oscillations of the shock. As
such, shock smoothing is employed to better resolve the flow disconti-
nuity. This is performed in two steps. The original steady RANS flowfield
is first smoothed across the entirety of the domain, producing the
smoothed field qsmooth . The smoothed and original fields are then
blended through:

qfinal ¼ ð1  SÞqoriginal þ Sqfinal (9)

where S is a blending function.


The discretised form of the eigenvalue problem posed by Equation (8)
is then solved through the implicitly restarted Arnoldi method [25].
Calculations are concentrated about the least stable eigenvalue, which is
tracked as incidence is increased. As the growth rate becomes positive,
instability and flow unsteadiness ensue.

2.2.3. Validation of predictions


The use of global stability for the prediction of flow unsteadiness due
to transonic shock oscillation was first applied by Crouch et al. [21] to the
NACA 0012 aerofoil, for which earlier experiments were conducted by
McDevitt and Okuno [26]. The authors considered the high Reynolds
(Re ¼ 10  106 ) data set with M ¼ 0:76. At these experimental condi-
tions, shock buffet onset occurred at α≈3 . As evident in Fig. 3, an in-
crease in the angle of incidence results in the least stable eigenvalue
crossing the real axis at α ¼ 3:03∘ , indicative of instability due to a Hopf
bifurcation. Additionally, as the incidence increases further, so too does
the instability growth rate.
In Fig. 4, the unstable u-velocity mode shape computed by Crouch
et al. [21] is given. As evident in the magnitude plot of Fig. 4(a), un-
steadiness is concentrated at the shock location and in the shear layer
downstream of the shock. The phase plot provided in Fig. 4(b) indicates
phase-locked motion of the shock and separated boundary layer - the
shear layer thins as the shock travels downstream and thickens during
upstream shock excursions. This qualitative behaviour is in good agree-
ment with the observations made by McDevitt & Okuno [26].
The authors further investigated the influence of shock resolution and
the number of shock smoothing cycles performed. The onset of instability
was found to be insensitive to both parameters, with similar predictions
of the critical incidence across each level of grid refinement and number

Fig. 4. u-velocity magnitude and phase for the unsteady-mode eigenfunction. NACA 0012
aerofoil results at the conditions: Re ¼ 10  106 , M ¼ 0:76 & α ¼ 3:2 (Crouch et al. [21]).

of smoothing cycles. However, an increase in shock thickness (resulting


from either coarser grid resolution or an increased number of smoothing
cycles) did have a pronounced effect on the critical frequency.
In a further study, Crouch et al. [10] explored the origins of transonic
shock buffet on the NACA 0012 through global stability analysis. In
Fig. 5, the buffet onset boundary computed by the authors through
URANS simulations and global stability theory are compared to the
experimental results by McDevitt & Okuno [26]. The results are in
excellent agreement up to M ¼ 0:8, supporting the global mode insta-
bility as the underlying mechanism for transonic buffet. Some discrep-
ancies do appear at M ¼ 0:8, however as both the global stability and
URANS results predict steady solutions at this flow condition, the poor
performance here is likely due to deficiencies in the calculations of the
steady flowfield.
The analysis provided by Crouch et al. [10] also gives evidence
against two classical indicators of buffet onset. The Mach number just
ahead of the shock at the various conditions considered does not appear
Fig. 3. Least stable eigenvalues of the NACA 0012 at Re ¼ 10  106 & M ¼ 0:76 for to yield a reliable indicator of buffet onset as postulated by Mabey [2].
various angles of attack (Crouch et al. [21]).

5
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

The unstable global mode mechanism is also supported by the


eigenvalue decomposition of the global Jacobian matrix performed by
Sartor et al. [29]. The authors find that the majority of system eigen-
values are independent of the angle of incidence, excluding the eigen-
value that lies in the vicinity of the buffet frequency. As evident in Fig. 7,
the least stable eigenvalue migrates towards the stability boundary with
increasing incidence (indicated by the dashed line), with instability
resulting from a Hopf bifurcation. A further increase in angle of attack
causes this eigenvalue to exhibit an inflexion, returning to the stable
region once more at buffet offset. The authors also identify that in
addition to the low-frequency shock oscillations associated with shock
buffet and the globally unstable mode, a medium frequency scale un-
steadiness may be present in the separated shear layer. While this is not
linked to the onset of buffet, it is indicative of the additional presence of a
Kelvin-Helmholtz type instability that is broadband in nature.

2.3. Transonic pre-stall instability

Fig. 5. Buffet onset boundary for an NACA 0012 aerofoil, with URANS simulation results In a computational investigation considering a number of aerofoils,
and experimental data of McDevitt & Okuno [26]; Re ¼ 10  106 (Crouch et al. [10]).
Iovnovich & Raveh [30] identified a number of characteristics common
to the buffet cycles of various wing sections. The authors classify shock
Further, comparisons of upper surface skin-friction coefficient at various buffet as a transonic pre-stall instability consistent with the unstable
flow conditions in the vicinity of the buffet boundary are in contradiction shock-wave/separation bubble interaction described by Raghuna-
with Pearcey's [6] bubble bursting hypothesis. At M ¼ 0:72, the separation than [14].
bubble induced by the shock is present at both pre- and post-buffet flow Iovnovich & Raveh employed the finite difference Riemann Elastic
conditions, whereas at M ¼ 0:80, the flow is completely separated in both Zonal Navier-Stokes Solver (EZNSS) [31], using the Spalart-Allmaras
instances, and yet, steady flow is predicted. Such findings show no cor- [32] turbulence model with the Edwards and Chandra [33] correction
relation between the bursting of separation bubbles and the onset of flow to study shock buffet cycles of the NACA 0012, RA16SC1 and NACA
unsteadiness. 64A204 aerofoils. The authors found that the range of incidence for
The shock motion detailed by Crouch et al. [10] provides a conflicting which the buffet instability is observed is narrower for the thin NACA
description of the transonic buffet phenomenon to the model posed by 64A204 aerofoil than the thicker sections. Nonetheless, a common
Lee [13]. While both models exhibit phase-locked modulation of the feature noted across each aerofoil examined was that the onset of global
separated shear layer and shock location, the model posed by Crouch unsteadiness occurred once the shock location moved aft of the upper
et al. [21] indicates pressure perturbations originating from the shock surface position of maximum curvature.
foot travel in the wall normal direction along the shock, and with lesser Examining the Mach number contours and flow streamlines for each
intensity, through the boundary layer, rather than simply through the of the aerofoils, they conveyed the common qualitative flow features of
shear layer. As this perturbation strikes the top of the shock it propagates the unstable shock/separation bubble interactions driving the shock
forward and dissipates into the oncoming flow. This behaviour is evident motion. Typical contours taken from the NACA 0012 are reproduced in
in the pressure fluctuations given at various steps across a shock cycle in Fig. 8. The cycle begins with the shock at its most downstream location,
Fig. 6. The pressure perturbations also move aft as they travel along the where it interacts with the separation bubble at the shock foot in
shock, intensifying during the downstream excursion. The acoustic Fig. 8(a). This interaction is unstable, with the high pressure bubble
waves then travel around the trailing edge and propagate upstream along pushing the shock upstream in Fig. 8(b).
the lower surface. Analogous findings are also presented by the same The upstream excursion of the shock is initially accompanied by an
authors for the OAT15A aerofoil [22], where comparisons between increase in shock strength, and hence, thickening of the separated shear
transverse velocity components from the global mode computations and layer. This is contrary to expectations from steady flow results, where an
Laser Doppler Velocimetry (LDV) measurements from experiment are in upstream shock position is typified by a reduced shock strength. The
excellent agreement. authors identify three factors that contribute to this shock strengthening:

2.2.4. Subsequent studies 1. Wedge Effects - Pressure rise due to the shock results in flow sepa-
Following the global mode instability proposed by Crouch et al. [21], ration emanating from the shock foot. This separated region behaves
a number of authors have reported findings that are in support of this similar to a geometric wedge, strengthening the shock, as evident in
interpretation. As discussed in Section 2.1, the modified wave- the increase in the oblique shock inclination angle between Fig. 8(a)
propagation model proposed by Jacquin et al. [20] encompasses pre- and (b).
cisely the qualitative flow features described by Crouch et al. [10]. 2. Dynamic Effects - As the shock moves upstream the velocity of the
Studies by Kuzmin & Shilkin [27] and Kuzmin [28] also draw links be- shock increases the relative Mach number of the upstream flow. With
tween buffet onset and a global mode instability, finding unsteadiness to shock strength related to M 2 , a strengthening of the shock is to be
be a consequence of a supercritical Hopf bifurcation of the global flow- expected.
field. The authors also note non-uniqueness of the global flowfield, where 3. Aerofoil Curvature Effects - The expansion of flow through the sonic
realisation of a definite solution is dependent on the time histories of the zone is dependent on the local surface curvature. The greater the local
flow conditions. Studying the fixed-point stability of the BAC 3-11/RES/ curvature of the aerofoil, the more pronounced the reduction in shock
30/21 aerofoil at pre-buffet conditions, Nitzsche [9] found that, in a strength.
linear sense, shock oscillations may be attributed to a natural resonance
in the steady transonic flowfield. With incidence increasing towards As the downstream location of the shock is near the location of
buffet onset, small perturbations in pitch, flap and longitudinal trans- maximum upper surface curvature, the curvature effects are expected to
lation show a reduction in damping, indicative of a global mode be small relative to the wedge and dynamic effects, resulting in a net
approaching instability. shock strengthening. As the shock achieves its most upstream excursion

6
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 6. Contours of the pressure fluctuation at eight steps during the oscillation cycle (Crouch et al. [10]).

in Fig. 8(c), curvature effects are more pronounced, weakening the shock the cycle repeating as it reaches its most downstream position.
strength and allowing reattachment at the shock foot. It is noted that the Analogous to Crouch et al. [10], Iovnovich & Raveh [30] assessed the
time lag between shock weakening and complete reattachment of the influence of separation bubble bursting on the onset of shock buffet. In
separated flow is likely a result of pressure wave propagation times be- Fig. 9, the resultant steady skin friction coefficients for each of the
tween the shock and trailing edge. As the shear layer reattaches, the aerofoils at buffet onset and 0:1 below the onset incidence are shown.
shock strengthens as it makes a downstream excursion in Fig. 8(d), with Noticeably, both the RA16SC1 and the NACA 64A204 aerofoils show a

7
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 7. Change in least stable eigenvalue with angle of attack at M ¼ 0:73 (adapted from
Sartor et al. [29]).

burst separation bubble both at, and prior to, onset. Conversely, the
NACA 0012 steady flowfield exhibits a separation bubble at pre- and
post-buffet conditions. With no apparent change in the nature of the skin
friction coefficient as unsteady shock motion develops, this investigation
presents further evidence against the bubble bursting theory of Pear-
cey [6].

3. Numerical investigations of transonic shock oscillations on


aerofoils

3.1. Reynolds-averaged Navier-Stokes simulations

The intricate flow features associated with the transonic buffet phe-
nomenon suggests the need for computationally taxing scale-resolving
simulations to model the instability. Nonetheless, a plethora of numeri-
cal studies have been devoted to assessing the efficacy of URANS
methods in capturing shock buffet [34–36]. Additionally, many such
authors have reported good correlations to experiment of the bulk flow
features under a Reynolds-averaged formulation [29,30,37,38]. The low-
frequency shock motion that is characteristic of the transonic buffet
phenomenon provides an explanation for the success of the URANS
approach. A number of authors [10,29] have noted that the global flow
unsteadiness exists on timescales orders of magnitude longer than those
of the shear layer eddies. As such, while the inherent averaging process of
URANS simulation is not able to resolve turbulence of varying scales,
fundamental buffeting flow features can be predicted with a fair degree
of accuracy. Nonetheless, such URANS simulations do exhibit a high
sensitivity to various simulation parameters, particularly the turbulence
model, spatial and temporal discretisation and the numerical scheme
used. The following section is devoted to highlighting the methodologies Fig. 8. Shock buffet cycle Mach number and flow streamline snapshots for the NACA
that have been most successful in reproducing transonic shock 0012 aerofoil at developed buffet conditions, M ¼ 0:72, α ¼ 6 (Iovnovich & Raveh [30]).
oscillations.
viscosity model at capturing transonic shock oscillations. The authors
3.1.1. Effect of turbulence model found that only the Spalart-Allmaras and NLEVMs with functional eddy
URANS simulation of shock buffet phenomena has been shown to be viscosity coefficients developed shock unsteadiness comparable to
particularly sensitive to the choice of turbulence model. Barakos & Dri- experiment, albeit, at a higher incidence and Mach number.
kakis [39] explored the effectiveness of various linear and nonlinear eddy In two broad spanning studies, Goncalves et al. [47] and Goncalves &
viscosity models (NLEVM) using an implicit unfactored Riemann solver Houdeville [37] employed the implicit, cell-centred finite volume CAN-
of third-order spatial and second-order temporal accuracy [40,41]. The ARI code [48] to assess the influence of various numerical parameters on
authors considered transonic flow over the NACA 0012 aerofoil, taken buffet predictions for the RA16SC1 aerofoil. The authors considered an
from the experiments of McDevitt & Okuno [26] to evaluate the effec- array of two-equation turbulence closures, including the Smith k l
tiveness of the Baldwin-Lomax [42], Spalart-Allmaras [32], Launder- [49,50], Wilcox k ω [51], Menter SST k ω [52], Kok k ω [53] and the
Sharma [43] and Nagano-Kim [44] linear k ε models and the high Reynolds variant of the Jones-Launder k ε [54] models, in addition
Sofialidis-Prinos [45] k ω version of the Craft et al. [46] nonlinear eddy- to the one-equation Spalart-Allmaras [32] model. All turbulent

8
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 9. Static friction-coefficient variation along the upper surface at pre-onset and onset conditions for three aerofoils (Iovnovich & Raveh [30]).

convective quantities are discretised through a second-order accurate


upwind Roe [23] scheme with flux-limited dissipation and Harten's en-
ergy correction [55]. Additionally, for each turbulence model the authors
applied an analytical velocity profile in the near wall region coupled with
the no-slip condition. The resulting predictions of buffet frequency and
lift differential for various angles of attack are provided in Table 1. The
results indicated that Menter's SST model produces the best correlations
to experiment for all conditions, including the predictions of RMS pres-
sures provided in Fig. 10.
Analogous to the findings of Barakos & Drikakis [39], the k ε and k

Table 1
Frequency and amplitude of the lift coefficient with various turbulence models (Goncalves
& Houdeville [37]).

Model α ¼ 3 α ¼ 4 α ¼ 5

f (Hz) ΔCL f (Hz) ΔCL f (Hz) ΔCL

Experiment 88 0.11 100 0.308 Steady State


SA 82 0.0146 92 0.325 100 0.55
k l – – Steady State – –
k l corrected – – Steady State – –
k l SST 79.5 0.0084 97.6 0.296 101.8 0.53
k ε Steady State 95.6 0.17 97.6 0.43
Fig. 10. RMS pressure fluctuations over the RA16SC1 aerofoil - α ¼ 4 (Goncalves &
k ε SST Steady State 95.6 0.48 101.8 0.67
Houdeville [37]).
k ε Durbin 85.2 0.012 93.7 0.437 101.8 0.67
k ω Wilcox – – Steady State – –
k ω Menter – – Steady State – – l models perform poorly and predict very low levels of unsteadiness at the
k ω SST Menter 90 0.11 96.6 0.33 Steady State onset condition. The authors note that incorporating a realisability
k ω Kok Steady State 94.6 0.26 95.6 0.48
k ω SST Kok Steady State 94.6 0.26 96.6 0.445
correction does significantly improve their behaviour. The Spalart-
k ω Kok Durbin Steady State 94.6 0.26 96.6 0.45 Allmaras model is again able to reproduce the unsteadiness observed in

9
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

experiments; however, the buffet frequencies and amplitude are under- OAT15A aerofoil are used as a validation dataset for the turbulent clo-
stated. The Kok k ω yields unsteady flow, although onset in delayed to sures (M ¼ 0:73, α ¼ 3:5 ), with resulting buffet frequencies and lift
higher incidence and no noticeable effect of the SST correction or real- amplitudes reproduced in Figs. 12 and 13.
isability condition is observed. The Wilcox and Menter k ω models are Results pertaining to the scalar dissipation formulation (not shown)
unable to reproduce shock oscillation in the absence of the SST generally underestimate the buffet amplitude by a significant margin.
correction. With matrix dissipation employed, the LEA model performs poorly in all
To further examine the influence of the wall law approach on tran- instances, limiting its applicability in the prediction of shock buffet
sonic buffet computations, Thiery & Coustols [56] used the explicit cell- phenomena. The original form of the Spalart-Allmaras model also fails to
centred, finite volume code elsA [57] to simulate transonic flow over the produce unsteadiness on all grids, contrary to preceding studies
RA16SC1 and OAT15A aerofoils. A four-step Runge-Kutta scheme was [21,37,39,56]. Further, only the SST model with matrix dissipation on a
employed to achieve second-order accurate temporal resolution, in hybrid grid is accurately able to capture both buffet frequency and
addition to the Jameson scheme [58] with artificial dissipation for the amplitude. Subsequent analysis by the authors does reveal that this is,
inviscid fluxes and the Roe scheme [23] for the turbulent convective however, an artefact of insufficient resolution of the boundary layer. The
quantities, both of which achieved second-order spatial accuracy. Thiery transition from hexahedral to tetrahedral cells occurs in the near-wall
& Coustols [56] also assessed the Spalart-Allmaras and Menter's SST region within the developing shear layer, resulting in a form of numeri-
models, along with the k ϕ model [59] developed at ONERA. The au- cal buffet that does not permit a robust method of solution. Consistent
thors further performed simulations for the Spalart-Allmaras and Men- buffet predictions are provided in the form of the εh -RSM model, as
ter's SST model using two grids. The first grid is refined through the evident in Fig. 13. The Reynolds stress model is insensitive to grid type
boundary layer region to achieve an average y þ ¼ 1:1 in the shock region and the form of dissipation, producing excellent correlations in buffet
(denoted RG). The second grid was constructed by removal of the first 20 frequency, amplitude and RMS pressure fluctuations in the shock inter-
nodes in the wall normal direction of the refined grid, such that an action region and at the trailing edge.
analytical velocity profile may be applied in this near wall region In addition to the studies dedicated to assessing the influence of
(denoted WL). turbulent closures, a number of authors have reported various degrees of
The influence of the wall law approach is considered on the RA16SC1 success with additional models. Kourta et al. [65] found a time depen-
aerofoil at M ¼ 0:732 and Re ¼ 4:2  106 . RMS pressure fluctuations are dent k ε model [66], where the model coefficient Cμ is related to local
well predicted for the majority of the models, excluding the RG SST deformation and strain rates, performed well in buffet computations of
formulation for which a steady solution was produced. The RG variant of the OAT15A aerofoil. The model is similar to the Zhu-Shih-Lumley model
the Spalart-Allmaras model does, however, appear to produce the most [67] employed by Brunet [68], who saw similar success. Soda & Verdon
consistent results, best approximating the RMS pressure fluctuations and [69] were able to capture shock oscillations over the NACA 0012 using
the lift amplitudes across a range of angles of attack, as shown in Fig. 11. the LEA [63] model, however, the Spalart-Allmaras closure with upwind
This indicates that while the WL approaches considered by Goncalves differencing in this study exhibited greater consistency across a range of
et al. [47] and Goncalves & Houdeville [37] may be able to reproduce the Mach numbers.
buffet phenomenon, better accuracy can be achieved with refinement Xiao et al. [17] found the lagged k ω model proposed by Olsen and
through the boundary layer. Coakley [70] produced good predictions of the mean pressure distribu-
Several common turbulence models were also assessed by Illi et al. tions of the BGK No. 1 aerofoil at buffet conditions, however, the
[60] using the unstructured finite volume code DLR-TAU [61]. The au- magnitude of oscillations was significantly overestimated through the
thors investigated the influence of matrix and scalar dissipation, com- shock region. Hasan & Alam [71] and Rokoni & Hasan [72] saw excellent
bined with the baseline Spalart-Allmaras model and the Strain Adaptive results in mean and RMS pressure distributions for the SC(2)-0714
variant (SALSA) [62], Menter's SST closure, the Linear Explicit Algebraic aerofoil using Menter's SST model. Carresse et al. [73] and Giannelis &
k ω model (LEA) [63] and the less commonly employed εh Reynolds Vio [74] also found Menter's SST model to perform best with the OAT15A
stress model (εh -RSM) [64]. The experiments of Jacquin et al. [20] for the aerofoil. These two studies further explored the efficacy of the RSM-stress
omega model (derived by coupling of the ω-equations with the Launder-

Fig. 12. Frequency and amplitude of shock motion for various turbulence models and
Fig. 11. Amplitude of lift coefficient for different turbulence models and the two RG and grids with matrix dissipation (S ¼ structured, H ¼ hybrid, G ¼ grid refinement level)
WL strategies (Thiery & Coustols [56]). (adapted from Illi et al. [60]).

10
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

been achieved with the Roe [28,38,73,74], Jameson [56,69,78,79] and


variants of the AUSM [18,37,69] schemes.

3.1.3. Effect of spatial & temporal discretisation


Several investigations have further considered the dependence of
buffet calculations on the temporal formulation employed [80,81].
Rouzaud et al. [80] investigated the effectiveness of an implicit Dual
Time Stepping (DTS) method in buffet computations for both an 18%
thick circular arc aerofoil and the RA16SC1 supercritical aerofoil. The
DTS method was found to produce results in fair agreement with
experiment and with substantial improvements in efficiency relative to
an explicit temporal formulation. These findings are in agreement with
the previous study of Rumsey et al. [81] for the 18% thick circular arc
aerofoil, where the use of subiterative DTS formulations saw improve-
ments in both temporal accuracy and convergence rates. Due to these
benefits offered by DTS, this implicit temporal formulation has been
employed extensively in recent URANS investigations on transonic buffet
[28,37,56,65,72,79].
The study by Rumsey et al. [81] further explored the influence of
Fig. 13. Frequency and amplitude of shock motion for the εh -RSM model (S ¼ structured,
temporal and spatial resolutions on predicted buffet characteristics. For
H ¼ hybrid, G ¼ grid refinement level, s ¼ scalar dissipation, m ¼ matrix dissipation, solutions performed with a subiterative DTS method, the authors found
δ ¼ height of hexahedral cells in original hybrid grid) (adapted from Illi et al. [60]). that with a temporal discretisation of approximately 170 steps per period
a 1% difference in computed buffet frequency was seen when the reso-
Reece-Rodi [75] model), which was capable of reproducing the buffet lution in time was increased by a factor of four. Conversely, significant
phenomenon, albeit with overestimated levels of unsteadiness in variations in the predicted buffet frequency and lift amplitude were
each case. observed when doubling the grid density, with differences of 4.5% and
6.5% respectively. Variation in the computed buffet characteristics of the
3.1.2. Effect of numerical discretisation scheme OAT15A aerofoil of similar magnitude with double the grid resolution
In addition to the influence of turbulence modelling, certain studies were also observed by Illi et al. [60]. Iovnovich & Raveh [30] further
have evaluated the significance of the numerical discretisation scheme highlighted the significance of sufficient shock resolution when per-
for the convective fluxes [37,69]. Goncalves & Houdeville [37] explicitly forming grid convergence studies for buffeting flows, with 10% and 12%
assessed the effects of the chosen numerical scheme for the mean flow- differences in buffet frequency and lift differential respectively as the
field on buffet predictions. The authors considered the Jameson [58], shock resolution was increased from 1.5% chord to 0.3% chord. The use
upwind Roe [23] with Monotone Upstream-Centred Scheme for Con- of a highly resolved grid in the shock region becomes particularly sig-
servation Laws (MUSCL) [76] extrapolation and the Advection Upstream nificant when using the linearised global stability method of Crouch et al.
Splitting Method (AUSMþ) [77] with MUSCL extrapolation schemes. Of [21] for buffet predictions. Shock resolutions on the order of 0.15%
the considered formulations, the Jameson scheme was found to be most chord were deemed necessary to accurately capture the flow physics.
effective in capturing onset, however the Roe scheme with MUSCL
extrapolation produced the best correlations to experiment regarding 3.1.4. Effect of wind tunnel geometry
shock frequency and lift differential. The resulting buffet frequencies and With the majority of URANS studies investigating the transonic buffet
amplitudes are reproduced in Table 2. The authors note that the influence phenomenon employing farfield boundary conditions, a select few au-
of the convective flux formulation on the predicted buffet characteristics thors have explored the influence of test section geometry on the
is secondary to the choice of turbulence model, with the various schemes computational results. Furlano et al. [82] performed an initial assessment
yielding a 3% difference in buffet frequency and 10% difference in lift of wind tunnel wall effects in steady conditions, using the ONERA [83]
differential. code to simulate transonic flow over the OALT25 and RA16SC1 aerofoils.
Soda & Verdon [69] also made a comparison of transonic flow pre- Two-dimensional steady results with various turbulence models offered
dictions made by upwind (AUSMþ) [77] and central differencing consistent results, with fair agreement to experiment for the OALT25
(Jameson) schemes of the inviscid fluxes for the NACA 64A010. The aerofoil. However, differences were observed in the pressure gradient
upwind scheme was found to improve the accuracy of the shock location, upstream of the shock and the local Mach number on the pressure side of
whereas the central scheme was more effective at capturing trailing edge the aerofoil. The authors attribute these discrepancies to three-
pressures. From the plethora of URANS investigations on transonic shock dimensional effects. Performing analogous simulations with the inclu-
buffet, it is evident that the conclusions of Goncalves & Houdeville [37], sion of the entire three-dimensional tunnel geometry produced better
regarding numerical discretisation as a secondary consideration relative predictions of the upstream pressure gradient and shock location. Gar-
to turbulence modelling, is well founded. With appropriate choice of baruk et al. [84] drew similar conclusions for steady transonic flow over
turbulence closure, successful reproduction of the buffet instability has the RAE 2822 aerofoil, stressing the need for inclusion of appropriate

Table 2
Frequency and amplitude of the lift coefficient with various discretisation schemes (Goncalves & Houdeville [37]).

Model α ¼ 3 α ¼ 4 α ¼ 5

f (Hz) ΔCL f (Hz) ΔCL f (Hz) ΔCL

Experiment 88 0.11 100 0.308 Steady State


Jameson 90 0.11 96.6 0.33 Steady State
Roe MUSCL 90 0.014 99.7 0.3 Steady State
AUSM þ MUSCL 90 0.018 98.6 0.307 Steady State
Jameson corrected 91 0.097 97.6 0.327 99.7 0.46

11
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

wind tunnel conditions during simulation. computational predictions of transonic shock buffet.
Thiery & Coustols [56] explicitly studied the influence of test section
geometry on the buffet predictions for the OAT15A aerofoil. An uncon-
fined farfield domain is compared with the true test section geometry, 3.2. Scale-resolving simulations
including adaptable tunnel walls, with time-averaged and RMS pressure
fluctuations reproduced in Fig. 14. The inclusion of test section geometry Although URANS simulation is effective in capturing the low-
yields a pronounced improvement in both the mean pressure distribution frequency shock oscillations that characterise transonic buffet, there
and RMS pressure fluctuations for the OAT15A aerofoil. The authors also remain several facets of these complex flowfields that simply cannot be
find that the predicted buffet frequency improves markedly, reducing to reproduced. In particular, these methods are unable to provide insight
74 Hz (relative to the experimental value of 69 Hz). Similar improve- into the effects of the broadband turbulent spectrum that inherently re-
ments are noted for the non-dimensional velocity profiles at various sults from the intermittently separated flow. Nonetheless, the past decade
chordwise stations. Barbut et al. [85] and Braza [86] also found the ef- has seen substantial developments in unsteady hybrid RANS/Large Eddy
fects of wind tunnel walls to be significant in the prediction of shock Simulation (LES) methods that resolve, rather than model, these turbu-
oscillations on the NACA 0012. The absence of the walls resulted in lower lent scales. In combination with the continuing growth of computational
levels of surface pressures and predicted onset at higher Mach numbers power, these scale-resolving simulations are steadily becoming a feasible
and incidence. The pronounced influence of the test section geometry on tool to probe the physics governing transonic buffet.
the transonic flowfield demonstrated by these preceding studies suggests One of the first investigations of transonic buffet over an aerofoil
that the inclusion of wind tunnel walls may yield improvements in through scale-resolving methods was performed by Deck [18]. Using the
FLU3M code of ONERA, Deck [18] compared URANS, Detached-Eddy
Simulation (DES) and the novel Zonal DES (ZDES) predictions of peri-
odic shock motions over the OAT15A aerofoil. A second-order accurate
upwind finite volume discretisation of the Navier-Stokes equations is
employed, with the upwind AUSMþ(P) [87] scheme used to resolve the
inviscid flux terms. A second-order accurate implicit temporal dis-
cretisation derived from Gear's formulation [88] is also used for the
unsteady simulations. Closure of the Navier-Stokes equations in the
URANS computations is achieved with the Spalart-Allmaras [32] model.
The DES calculations follow the original formulation developed by Spa-
lart et al. [89] and are again based on the one-equation Spalart-Allmaras
[32] model. The reader is directed to the original work for a compre-
hensive derivation of the DES formulation, however, the critical aspect
outlined by Deck [18] is the inclusion of an eddy viscosity destruction
term that is a function of the distance to the nearest wall (d). In
conjunction with the production term, the computed eddy viscosity (~ν)
~ such that:
scales with the local deformation rate (S)

~ν≈~Sd2 (10)
Spalart et al. [89] suggested replacing the original length scale d with
a modified scale:

d~ ¼ minðd; CDES ΔÞ (11)

where Δ ¼ maxðΔx ; Δy ; Δz Þ represents the grid size and CDES is a model


constant. The dependence of the length scale on the maximum grid
extension is a natural representation, as this distance governs the
wavelengths that can be resolved by the simulations.
While the standard DES formulation seeks to combine the best attri-
butes of both RANS and LES simulations, the application to transonic
buffet, where intermittent thin-layer separation is present, may be
problematic. In particular, premature switching to LES mode may occur
within the RANS boundary layer, yielding fictitious grid-induced sepa-
ration. In an attempt to remedy this, Deck [18] developed the ZDES
method. In this approach, the user explicitly identifies the RANS and LES
regions at the outset of the simulation. The attached shear layer region
and the entire shock/boundary layer interaction zone are treated exclu-
sively in RANS mode, while the separated flow region aft of the trailing
edge is resolved in LES.
Deck [18] further took advantage of an innovative two-three-
dimensional grid coupling technique. The method was developed by
Mary & Sagaut [87,90] under the LESFOIL project to reduce the high-
density grid resolution inherent in scale-resolving simulations. In
Fig. 15, a typical grid topology for this coupling method is shown.
Notably, only zones 3 and 6 are treated in LES. Further, three-
dimensional resolution is limited to zones near the aerofoil and in the
Fig. 14. Effect of wind tunnel walls on the unsteady pressure distributions of the OAT15A wake, with the remainder of the domain computed in two-dimensions,
aerofoil (α ¼ 3:5 , M ¼ 0:73, Re ¼ 3  106 ) (Thiery & Coustols [56]). reducing the node count by a factor of two relative to an entirely

12
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

longitudinal velocity profiles and fluctuations, as evident in Fig. 17.


Discrepancies are evident in these profiles upstream of the shock; how-
ever, the author attributes this to the calculations being performed in a
fully turbulent manner, whereas the experiment employed a fixed tran-
sition trip. Power spectral analysis of the wall pressures, constructed
through a parametric autoregressive technique [91] due to the limited
available simulation time, indicate the dominant buffet frequency is well
captured by ZDES at the experimental condition and by URANS and DES
at higher incidence. Flattened spectra at higher frequencies, an indication
of random white noise associated with varying scales of turbulent
structures, are also best captured by the ZDES simulations.
Analogous to Xiao et al. [17], Deck [18] assessed the efficacy of Lee's
[13] model at predicting the buffet frequency of the OAT15A aerofoil.
Using a frequency-wave number spectrum [92] to compute the down-
stream wave propagation velocity, the computed buffet frequency
derived from the computations is in excellent agreement with the
experiment, differing by approximately 6%.
Although this initial study appears to support Lee's [13] model, a
subsequent investigation by Garnier & Deck [19] contradicts the early
Fig. 15. Two-three-dimensional grid (Deck [18]). findings. Again using the second-order accurate in space and time FLU3M
solver, the authors employed a modified Roe scheme [23] for the
three-dimensional grid. convective fluxes, integrating the Ducros et al. [93] sensor to locally
The resultant calculations clearly indicate that the ZDES method is adapt the solvers dissipation. A zonal RANS/LES method is again
superior to both URANS and classical DES in terms of mean and RMS employed to resolve the turbulent eddies, however, the entirety of the
pressure fluctuations and general character of the oscillating flowfield suction side and wake region are now treated through LES, with the
(evident in the RMS of longitudinal velocity fluctuations in Fig. 16). pressure side resolved in two-dimensional RANS mode. For sub-grid scale
Notably, the ZDES results overestimate the degree of pressure fluctuation modelling, the Selective Mixed Scales Model (SMSM) [94] has been used.
both through the shock region and at the trailing edge, as well as pre- Investigating the influence of spanwise grid extent, Garnier & Deck
dicting a shock location marginally upstream of the experiment. Further, [19] found a two-fold increase in the domain size (from 3.65% chord to
while the URANS and DES simulations require a higher angle of inci- 7.30% chord) significantly reduced the computed pressure fluctuations
dence relative to the experiment before unsteadiness is observed (α ¼ at the trailing edge. As shown in Fig. 18, the larger transverse grid extent
4:5 and α ¼ 4 respectively), ZDES is able to capture shock oscillation at (B1 and B2) better captures the unsteady pressures at the trailing edge by
the experimental incidence of α ¼ 3:5 . allowing three-dimensional coherent structures to develop, limiting their
The ZDES simulations also provide good predictions of downstream intensity relative to the predominantly two-dimensional structures

Fig. 16. Longitudinal velocity fluctuation, RMS field (Deck [18]).

13
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 18. RMS pressure distribution for the OAT15A (adapted from Garnier & Deck [19]).

the near-field flow to compute wave convection velocities, Garnier &


Deck [19] also assessed the accuracy of Lee's [13] model for buffet fre-
quency prediction. A significant discrepancy is apparent between the LES
buffet frequency of 72 Hz and the analytical prediction of 110 Hz. The
authors do note, however, an upstream propagating pressure disturbance
along the lower surface of the aerofoil, consistent with the global mode
instability of Crouch et al. [21].
Furthering Deck's [18] study, Grossi et al. [38] compared the effec-
tiveness of URANS and Delayed DES (DDES) simulations in the prediction
of transonic buffet for the OAT15A section. The authors employed the
density based finite volume Navier Stokes Multi Block (NSMB) code [95],
with the third-order accurate total variation diminishing (TVD) variant of
the upwind Roe scheme [23] with MUSCL extrapolation for the
convective fluxes and a second-order accurate backward Euler dual time
stepping scheme for temporal discretisation. For the two-dimensional
URANS computations a number of turbulence closures were investi-
gated, from which, the Edwards-Chandra modified Spalart-Allmaras
model with compressibility correction (EDW-CC) [33] was found to
best capture the flow features observed at equivalent conditions to the
experiments. The hybrid DDES formulation is developed from the un-
derlying EDW þ CC RANS turbulence model, and employs a modified
length scale relative to the standard DES formulation of Equation (11),
defined by:

d~ ¼ d  fd maxð0; d  CDES ΔÞ (12)


Fig. 17. Velocity profile predictions (Deck [18]).
where fd is a delaying function.
The analysis provided by Grossi et al. [38] indicates the formulation
observed with the original grid (A). Evidently, the intensity of pressure of turbulent quantities has a significant effect on the flow topology pre-
fluctuations is overestimated relative to the experiment in each instance. dicted. From the DDES computations, the separated flow immediately aft
Additionally, the averaged pressure distribution of Fig. 19 indicates a of the shock develops as a primarily two-dimensional Karman instability.
downstream mean shock location. The authors note that the level of As the shock moves upstream, a spanwise undulation develops. From this
numerical dissipation imparts an appreciable influence on the predicted perturbation, a secondary instability emanating from the primary
shock location. The high dissipation simulation (B1) yields a mean shock vortices ensues and the two-dimensional structures break down, exhib-
location further aft, with the converse applying in the low dissipation iting a strongly three-dimensional character downstream. Further, power
case (B2). spectral densities (PSDs) of the lift time history do not display sharply
Relative to the earlier work by Deck [18], the LES simulations per- resolved peaks, but rather, dispersed bumps. Such spread in the fre-
formed by Garnier & Deck [19] yield more consistent predictions of quency spectrum is indicative of an aperiodic and broadband buffet
longitudinal velocity profiles and fluctuations along the entire chord. As response. Conversely, the URANS simulations show a periodic character
shown in Fig. 20, regardless of the degree of numerical dissipation, the with no appearance of a secondary instability. The differences in flow
longitudinal velocity fluctuations are well captured. Spectral analysis of topology predicted by the URANS and DDES approaches can be seen in
the wall pressure fluctuations reveals a dominant buffet frequency of Fig. 21, which provides a map of the separated flow regions over a
72 Hz, consistent with the findings of Deck [18] and marginally higher buffet cycle.
than the experimental value of 69 Hz. Further discrepancies between URANS and DDES simulations are
Using two-point two-time correlations of the fluctuating pressures in

14
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 19. Mean pressure distribution for the OAT15A (adapted from Garnier & Deck [19]).

Fig. 21. Spatiotemporal evolution of flow separation on the upper surface (Grossi
et al. [97]).

present in the analysis of statistical flow properties, particularly in the


shock/boundary layer interaction region and at the trailing edge. The
DDES simulations predict with fair accuracy the shock location, range of
shock travel and peak pressure RMS in the shock region, however, the
pressure fluctuations at the trailing edge are significantly overestimated,
as found by Deck [18]. Conversely, although the URANS simulations
yield a shock location downstream relative to the experiment, mean and
RMS trailing edge pressures are well captured. Mean and RMS longitu-
dinal velocities are also best represented by the URANS approach. Such
results suggest that the preferred method of solution may be dependent
on the particular features of buffet to be investigated, with the DDES
approach providing a richer resolution of qualitative flow features and
Fig. 20. Longitudinal velocity RMS profiles (adapted from Garnier & Deck [19]). the URANS method better predicting the statistical properties.

15
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Grossi et al. [96] extended their previous study, performing both two-
and three-dimensional URANS simulations, in addition to three-
dimensional DDES with the SALSA [62] turbulence model and the
third-order AUSMþ [77] scheme for the convective fluxes. While refor-
mulation of the underlying RANS model in the DDES approach improves
the prediction of pressure fluctuations at the trailing edge and shifts the
shock location in the URANS method upstream, the findings are generally
in accordance to Grossi et al. [38]. The results of both papers are pre-
sented in greater detail in Grossi's dissertation [97].
As an alternative to the ZDES method presented by Deck [18], Roidl
et al. [98] developed a novel zonal hybrid RANS-LES method to compute
transonic shock oscillations over the DRA 2303 aerofoil [99]. The com-
putations employ a mixed second-order accurate centred/upwind AUSM
[100] scheme for spatial discretisation of the inviscid fluxes, with second-
order central differencing for viscous terms. Temporal discretisation also
achieves second-order accuracy through an explicit 5-stage Runge-Kutta
method. To resolve turbulent quantities, the domain is segmented as
shown in Fig. 22. The Spalart-Allmaras [32] model is used in the RANS
zone and the Monotone Integrated LES (MILES) [101] approach in the
LES region. The hybrid method proposed employs a synthetic turbulence
generation method (STGM) at the LES inlet and localised control planes
to modify turbulence production in a buffer region between the two
zones. Further details of the formulation are provided by Zhang
et al. [102].
Two distinct grids were developed for the computations; a pure LES
grid of resolution Δxþ ≈100, Δy þ ≈1 and Δzþ ≈20 for the streamwise, wall
normal and spanwise directions, and a hybrid grid of equivalent resolu-
tion in the LES zone and coarser streamwise and spanwise spacing in the
RANS region. Although no direct validation of either method relative to
experiment is provided, the hybrid and pure LES approaches yield
consistent predictions for statistical flow properties (Fig. 22), Reynolds
stresses (Fig. 23(b)) and pressure PSDs. The hybrid method is thus able to
achieve an equivalent degree of accuracy relative to pure LES with a
significant reduction in the grid requirements, in this instance, reducing
the node count by a factor of two.

4. Experimental investigations of transonic shock oscillations on


aerofoils

Recent years have seen few experimental investigations in open


literature that explicitly study the nature of transonic shock oscillations
on aerofoils. Many such studies were conducted following the seminal
Fig. 23. Validation of the Zonal RANS-LES method (M ¼ 0:72, Re ¼ 2:6  106 , α ¼ 3 )
work of Tijdeman [11] in classifying the various types of shock motion (Roidl et al. [98]).
observed on the NACA 64A006 excited by sinusoidal trailing edge flap
deflections. McDevitt et al. [103], McDevitt [104], Mabey [2] and Mabey Hirose & Miwa [109] and Lee [15,110], while McDevitt & Okuno [26]
et al. [4] each performed wind tunnel tests on biconvex aerofoils. Ex-
investigated the NACA 0012 buffet boundary. Although the volume of
periments considering shock oscillations over supercritical aerofoils were transonic buffet experiments in recent years has diminished, a number of
conducted by Finke [105], Stanewsky & Basler [106], Roos [107,108],
comprehensive studies have appeared which have been instrumental in

Fig. 22. Schematic of the zonal RANS-LES method for transonic flow around the DRA 2303 profile (Roidl et al. [98]).

16
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

furthering the understanding of the underlying flow physics. spectrogram in Fig. 25(a) and sound pressure level contours in Fig. 25(b),
An extensive computational and experimental campaign has been indicate that the two-dimensional buffet phenomenon is time-invariant
undertaken at ONERA over the past decade, seeking to characterise the and essentially modal in nature. Excluding the intermittent shock re-
buffet phenomenon and develop alleviation techniques. A complete gion in Fig. 25(b), where frequency spreading is present due to the low-
overview of the programme is given by Dandois et al. [111]. In particular, frequency shock oscillation, the spectra are dominated by a single fre-
experiments on the supercritical OAT15A performed in the S3Ch quency. Such behaviour is in support of the global mode instability
Continuous Research Wind Tunnel at the ONERA Chalais-Meudon Center theory proposed by Crouch et al. [21].
have been detailed by Jacquin et al. [20,112]. A wind tunnel model of In analysis of the two-dimensional character of the buffet phenome-
12.3% relative thickness, 230 mm chord, 780 mm span and a 1.15 mm non, Jacquin et al. [20] observed somewhat conflicting results. While
thick trailing edge was constructed for the experiment. The model spectral analysis revealed a constant spanwise distribution of sound
ensured a fixed boundary layer transition at 7% chord through the pressure levels across the central span, oil flow visualisations in the
installation of a carborundum strip on the upper and lower surfaces. separated region indicated the presence of three-dimensional structures.
The experiments conducted under this campaign sought to develop an The authors note that the velocity associated with the pressure fluctua-
extensive experimental database for the validation of numerical buffet tions is an order of magnitude larger than that of the three-dimensional
simulations. The model was fitted with 68 static pressure orifices and 36 structures observed in the wall velocity field. As such, Jacquin et al. [20]
unsteady Kulite pressure transducers through the central span, mitigating hypothesised the buffet instability may result from the superposition of
the three-dimensional effects from sidewall boundary layers. Adaptable strong two-dimensional and weaker three-dimensional global modes.
upper and lower wind tunnel walls (Fig. 24) further alleviated wall Jacquin et al. [20] also applied a similar line of reasoning to Lee [13],
interference, allowing for a test section Mach number uncertainty of that the observed buffet period for an aerofoil consists of the sum of
104 . A sublimating product was applied to the models upper surface, disturbance convection time-scales, to develop the modified wave-
which permitted oil flow visualisations for characterisation of turbulent propagation feedback model briefly detailed in Section 2.1. The con-
regions and shock motion. The authors employed Schlieren cinematog- stituent time-scales include the time taken for a disturbance originating
raphy to observe qualitative flow features and two-component Laser at the shock foot to propagate to the trailing edge (τp ) and an acoustic
Doppler Velocimetry (LDV) to capture quantitatively the longitudinal time-scale which defines the time delay between perturbations
and vertical velocity fluctuations during the buffet cycle. Further, steady emanating from the trailing edge impinging on the shock (τu ). The au-
and unsteady pressure measurements produced mean and RMS pressure thors note that reduction of the global physics governing the buffet
data, along with spectral content for the pressure fluctuations. phenomenon to a simple empirical model is subject to several difficulties,
The test programme undertaken by Jacquin et al. [20] consisted of an particularly in the computation of appropriate convection velocities.
angle of attack sweep at M ¼ 0:73 to obtain data for buffet onset, as well Whereas Lee [13] postulated τu represented the time taken for
as Mach number sweeps at α ¼ 3 and α ¼ 3:5 . The chord-based Rey- acoustic waves to convect upstream above the upper aerofoil surface
nolds number was kept approximately constant at Re≈3  106 for each from the trailing edge (τu;u ), Jacquin et al. [20] proposed such distur-
test point. The wealth of steady and time-resolved pressure and velocity bances could also propagate towards the shock along the lower surface,
field measurements produced during the programme have since seen the rounding the leading edge and impinging on the shock from upstream
OAT15A aerofoil become somewhat of a benchmark transonic buffet test (τu;l ). In the latter, the authors found τu ¼ τu;u þ τu;l produced the most
case, as evidenced by the multitude of numerical studies on this section in consistent results relative to experiment. Pressure fluctuations from the
the recent years [18,22,56,73,74]. Kulite transducers indicated disturbance propagation along both surfaces
The extensive analysis performed by Jacquin et al. [20] also served to and a predicted buffet frequency with the modified definition of τu in
provide novel insights of the physics governing transonic shock oscilla- better agreement with the experiment.
tions. Spectral analyses of the unsteady pressure signals, as shown in the Jacquin et al. [20] also detailed difficulties in the calculation of the
convection velocity ap in the computation of τp . Two-point cross corre-
lation of pressure fluctuations are employed by the authors to compute
ap , with comparable velocities obtained to those of Lee [13]. These dis-
turbances do not lend themselves to a simplified interpretation, and
different values for the convection speed would be obtained if these
perturbations were viewed in light of local stability theory and charac-
terised the propagation of a Kelvin-Helmholtz type instability.
The authors conclude that while the modified propagation model
better represents the flow characteristics observed in their experiments, a
robust and simple model of the phenomenon remains elusive. The shock
oscillations are driven by two cooperating mechanisms; propagation of
perturbations emanating from the shock foot in a region of receptivity
and forcing of these perturbations through the convection of acoustic
waves. The complexity of the constituent flow mechanisms, as they are
currently understood, does not permit the construction of a general-
ised model.
A contrary perspective has been provided by Hartmann et al. [113],
whose investigations support a second modified acoustic wave-
propagation mode. Experiments were conducted on the supercritical
laminar-type DRA 2303 aerofoil at Re≈2:7  106 , with a freestream
Mach number between 0:67  M  0:76 and incidence range of
0  α  4 . Steady and unsteady pressure measurements, time-resolved
stereo particle-image velocimetry (TR-SPIV) and Schlieren imaging are
used to characterise the unsteady flowfield across the range of flow
conditions. Additionally, the experiments are repeated at equivalent
Fig. 24. OAT15A supercritical profile in the S3Ch transonic wind tunnel (Jacquin conditions in the presence of an artificial acoustic source downstream of
et al. [20]).

17
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 25. Spectral analyses of OAT15A pressure fluctuations (Jacquin et al. [20]).

the test section. The artificial source is achieved through removal of the 2303 aerofoil, extending the original study to include a mid-range driver
upper and lower wind tunnel walls, with the freestream chamber acting and horn to allow for control of the acoustic driving frequency. This
to amplify acoustic disturbances at a frequency of 248 Hz. subsequent study details a single flow condition of M ¼ 0:73 and α ¼
The authors find that in the absence of the acoustic source, shock 3:5 . The baseline aerofoil exhibits Tijdeman Type A [11] shock motion
buffet is observed with M ¼ 0:72 and α ¼ 3 . The time series in Fig. 26 analogous to the cycle represented in Fig. 26, at a frequency of 129 Hz.
shows the velocity field and corresponding pressure distribution for a Two-point two-time cross correlations of both pressure and velocity
single buffet period at this condition. The shock oscillation is charac- signals are used to compute disturbance propagation velocities. Appli-
teristic of Tijdeman Type A [11] motion at a frequency of 126 Hz. cation of Lee's [115] model resulted in fair prediction of the buffet fre-
At equivalent flow conditions, the addition of an acoustic source has a quency (146 Hz), however, the authors suggest a modification to the
significant influence on the nature of the shock oscillations. PSDs of original model concerning the characteristic length for upstream acoustic
pressure fluctuations at the midchord indicate the frequency content is propagation. A change in the characteristic length from c xs to lshock =c in
dominated by the acoustic forcing frequency of 248 Hz; the acoustic Fig. 28 produced a significant improvement in the predicted frequency,
source has resulted in mode switching of shock oscillations to the first which correlates well with the experiment at 133 Hz.
harmonic of the undisturbed aerofoil. Additionally, time series of the Phase-averaged correlations of the streamwise and normal velocity
velocity fields in Fig. 27 show a dissimilar oscillation cycle relative to the fluctuations revealed highly correlated large-scale structures down-
Type A sinusoidal motion in Fig. 26. As the shock moves upstream from stream of the shock. Noting that a mechanism for noise generation is the
time ~t , the supersonic region contracts and the shock loses strength. At impinging of these vortical structures on the trailing edge [116,117],
t ¼ ~t þ 5Δt the primary shock region has practically vanished, and is Hartmann et al. [114] computed an acoustic disturbance of approxi-
accompanied by the formation of a weaker secondary shock, indicating mately 1030 Hz is emitted as the vortices pass the trailing edge. A pure
the beginning of a new cycle. Such an intermittent presence of the pri- sine wave at this frequency was then driven through the installed speaker
mary shock, weakening and disappearing during the upstream excursion, to amplify the effects of these high frequency perturbations. PSDs of the
cannot be characterised by any of the classical Tijdeman types. At higher shock location show frequency content at both the undisturbed buffet
Mach numbers, the intensity of the acoustic source is comparatively frequency and the acoustic driving frequency, with two additional modes
smaller, imparting less influence on the shock motion and producing appearing within this frequency band, indicating the shock motion is
lower amplitude oscillations. receptive to high frequency forcing.
The results presented by Hartmann et al. [113] categorically repre- Hartmann et al. [114] go on to provide a detailed analysis of the
sent transonic shock oscillations as an acoustic phenomenon. In the buffet mechanism strongly linked to noise generation and disturbance
absence of an external source, noise is produced at the trailing edge, with propagation. This proposed mechanism (for which a single oscillation
intensity varying dependent on the extent of the separated region. Larger cycle is shown in Fig. 29) is similar to that of Lee [13], consisting of a
separated regions, as seen at higher Mach numbers, result in reduced feedback loop between noise production at the trailing edge and the
sound pressure levels emanating from the trailing edge, and hence, a location of the shock. As the shock oscillates, the strength of vortices
reduction in unsteadiness. Through comparison of both forced and un- emanating from the shock foot and the extent of the separated region
forced experiments, the authors propose the noise generation at the vary. Both of these factors influence the distribution of the Lamb vector,
trailing edge drives the shock motion, which in turn, varies the extent of an indicator of trailing edge noise intensity [116,117]. Such changes in
the separated flow region and completes the feedback loop. the Lamb vector result in acoustic waves of varying sound pressure levels
Hartmann et al. [114] carried out further experiments on the DRA which emanate from the trailing edge. A decrease in sound pressure level

18
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 26. Time sequence of synchronously measured pressure distribution and velocity field for M ¼ 0:72, α ¼ 3 - no acoustic source (Hartmann et al. [113]).

results in a downstream shock excursion, and vice versa. motions, and the associated shock oscillations, on aerofoils at transonic
Although not explicitly considering large-scale shock motion, the conditions [120–122]. In an investigation involving transient pitching of
complex mechanisms governing the production of the trailing edge dis- a NACA 64A010 with shock-induced separation, Davis & Malcolm [123]
turbances detailed by Hartmann et al. [113,114] are investigated observed distinct resonance in the aerodynamic coefficients when the
experimentally by Alshabu et al. [118]. For the BAC3-11 aerofoil at M ¼ system was excited at a particular reduced frequency. The measured gain
0:71, Re ¼ 3:3  106 and α ¼ 0 , pressure fluctuations at the trailing edge and phase relationships are reminiscent of a pair of complex conjugate
gain strength as they propagate upstream, before interacting with the eigenvalues in a single degree-of-freedom harmonic oscillator. Similar
supersonic region and weakening. Unsteady pressure measurements resonant behaviour was also found by Despre et al. [124] during sinu-
indicate the dominant frequency of these trailing edge disturbances is in soidal trailing edge deflections of the OAT15A aerofoil in a buffeting
accordance with the findings of Hartmann et al. [114] and on the order of flow, where the maximum amplitude of shock-wave oscillation occurred
1 kHz. Further experiments conducted by Alshabu & Olivier [119] on the at excitation frequencies near the buffet frequency. As will be discussed,
BAC3-11 aerofoil showed that interaction between the shock and up- the prevalence of such an aerodynamic resonance near shock buffet onset
stream propagating disturbances resulted in degeneration of the shock has significant implications for the nature of shock oscillations in dy-
into compression waves coupled with the production of vortices in the namic aerofoil systems.
boundary layer. The vortices propagate toward the trailing edge, initi- In examining the fixed-point stability of the BAC 3-11/RES/30/21
ating a new cycle of trailing edge pressure waves in an analogous manner aerofoil at pre-buffet conditions, Nitzsche [9] also drew analogy between
to the mechanism detailed by Hartmann et al. [114]. shock buffet and a single degree-of-freedom oscillator. The DLR-TAU
[61] was employed with central differencing of the inviscid fluxes and
5. Dynamic interactions in the presence of transonic shock second-order backward Euler dual time stepping for temporal dis-
oscillations cretisation. Using the LEA variant of the k ω turbulence model [63],
buffet onset was captured by sweeping through incidence at M ¼ 0:75
Beginning with the original work of Tijdeman [11], a number of and Re ¼ 4:5  106 . Hysteresis was observed during the incidence
experimental studies have investigated the influence of forced harmonic sweeps, with buffet persisting to lower angles of attack during the down

19
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 27. Time sequence of synchronously measured pressure distribution and velocity field for M ¼ 0:72, α ¼ 3 - acoustic source (Hartmann et al. [113]).

locations in Fig. 30, where the resonant peaks align just below the onset
reduced frequency of ω ¼ 0:63.
From Fig. 30, the phase responses of each of the excitations are
consistent in the vicinity of the resonant peak. Specifically, the shock
buffet resonance is independent of the mode of excitation and is char-
acterised by a  180 phase reversal with a characteristic slope.
Considering the pitch phase response, Nitzsche [9] found that the shock
buffet eigenfrequency appears where the shock motion changes from
inverse to regular motion. Davis & Malcolm [123] encountered precisely
this behaviour in their transient pitching experiments on the
NACA 64A006.
In a comprehensive numerical study on the interaction between
forced harmonic heaving and shock buffet of the NACA 0012, Raveh
[125] uncovered a buffet lock-in mechanism related to the aerodynamic
resonance detailed by Nitzsche [9]. URANS simulations using the finite
difference code EZNSS [31] were performed with the modified k ω
turbulence model [126], and were validated against the static and dy-
namic experimental data of McDevitt & Okuno [26] and Landon [122],
respectively.
Fig. 28. Sound wave propagation and geometric flow properties (Hartmann et al. [114]).
Considering the nominal condition (M ¼ 0:72, Re ¼ 10  106 , α ¼
6 ), at which autonomous shock oscillations are observed for the rigid
sweep. For increasing angles of attack in the pre-buffet flowfield, fre- aerofoil, sinusoidal excitations in heave of varying frequency and
quency responses of the lift signal, excited by small oscillations of a amplitude were applied to the wing section. The resultant aerodynamic
trailing edge flap, developed more lightly damped characteristics and responses indicated that for sufficient amplitudes of motion and driving
pronounced resonant peaks at frequencies approaching the buffet fre- frequencies in the vicinity of the rigid buffet frequency, the buffet flow
quency. Similar behaviour was observed when varying the excitation response synchronises with the aerofoil motion. This behaviour is
kinematics to include perturbations in pitch and longitudinal translation, evident in Fig. 31, where the excitation amplitude increases, the fre-
as indicated by the frequency responses of the shock and separation point quency band for which lock-in occurs also broadens. This phenomenon is

20
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 29. Schematics of the flow characteristics for one oscillation cycle of the shock-wave motion, (a) when the shock is in its most downstream position, (b) shock-wave moves upstream at
its maximum velocity, (c) turning point of the cycle with the shock being in its most upstream position, (d) shock-wave moves downstream at its maximum velocity; the thickness of the
circles represents the strength of the vortices, ωI , ωII , ωIII , ωIV (Hartmann et al. [114]).

likened to flowfield of an oscillating cylinder, whereby the shedding Simulations of forced harmonic oscillations in heave reveal qualita-
frequency prevalent in the von Karman vortex street synchronises with tively different aerodynamic responses depending on whether the exci-
the driving frequency [127]. tation frequency is above or below the buffet frequency. At sufficient
At conditions which exhibited lock-in, Raveh [125] noted that for amplitudes for synchronisation to occur, driving frequencies above the
excitation frequencies below buffet, the lift response leads the aerofoil buffet fundamental result in quasi-harmonic oscillation of the aero-
motion. The converse is observed at higher driving frequencies, where dynamic coefficients, with smaller shock travel and less variation in
the lift response lags the motion. The lift responses were also seen to be shock strength. This is illustrated by the Mach contours in Fig. 34. The
symmetric about the rigid buffet mean lift and exhibited harmonic re- severe vortex shedding prevalent during the rigid shock oscillation cycle
sponses at frequencies above the rigid buffet fundamental. Such obser- of Fig. 33 has been quenched. This smoothing of the flowfield was also
vations allowed the extraction of a simplified gain-phase model, shown observed for small heave oscillations (equivalent to an induced incidence
for the lift-curve slope in Fig. 32. Analogous to Nitzsche [9], the lift-curve of 0:1 ) at the buffet frequency. At excitation frequencies below the
slope undergoes a phase reversal and exhibits a resonant peak at buffet, vortex shedding is again quenched, however an expansive sepa-
approximately the buffet frequency. Raveh [125] further posited as to rated region exists aft of the shock. The aerodynamic responses are no
how the lock-in phenomenon may provide a mechanism for LCO of longer harmonic and exhibit large fluctuations due to large-scale shock
aircraft in the transonic regime. Shock oscillations may serve to excite a motion. The authors note that qualitatively similar behaviour was
lightly damped structural mode during flight conditions near the stability observed for excitations in pitch. A recent investigation by Giannelis &
boundary, yielding increasing amplitude modal oscillations. With Vio [129] confirmed analogous flow features for a harmonically driven
increasing amplitudes lock-in occurs, and acts to reduce the excitation OAT15A supercritical aerofoil.
loads, with this cycle then repeating. Iovnovich & Raveh [130] continued to explore the lock-in mechanism
Raveh & Dowell [128] performed further numerical studies on the on the NACA 0012, with a particular focus on the aerodynamic response
lock-in mechanism for the NACA 0012 aerofoil under equivalent condi- due to pitch and flap excitations. At pre-buffet conditions, pitch excita-
tions, analysing the pre-buffet response to aerofoil motion and the in- tion of the aerofoil was found to yield a phase lead in the lift response for
fluence of lock-in on the flowfield topology. At pre-buffet conditions, the excitation frequencies below the buffet, with a phase reversal at the
flow was found to behave analogous to a linear oscillator, as per the buffet frequency. The converse was observed for both the moment co-
findings of Nitzsche [9]. As the mean angle of attack increases towards efficient for pitch excitations and the lift coefficient for flap excitations,
onset, so too does the resonant frequency of the flow, with a corre- where the responses exhibit a phase lead at higher driving frequencies.
sponding decrease in damping. Onset is characterised by zero damping in These characteristics persist in developed buffet flow for pitch excita-
the flow. At conditions in the region of developed buffet, the shock tions, and the combination of phase lead and significant aerodynamic
motion for the rigid aerofoil resembles Tijdeman [11] Type A oscilla- fluctuations support the transonic LCO mechanism posed by Raveh
tions. From the Mach number contours in Fig. 33, strong pressure fluc- [125]. The aerodynamic response due to flap motions is fundamentally
tuations at the trailing edge are observed during the shocks upstream different, acting to attenuate the buffet lift response and eliminate the
excursion (Fig. 33(d)), yielding prominent vortex shedding into the wake resonance lift response. These findings support the use of trailing edge
region (Fig. 33(e)–(g)). deflections as an effective means of shock buffet alleviation.

21
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 31. Combinations of excitation frequencies and angles of attack leading to responses
in both the excitation and buffet frequencies and in the excitation frequency only; α ¼ 6 ,
M ¼ 0:72, Re ¼ 10  106 (Raveh [125]).

excitation considered. This is illustrated by the aerofoil pitch and shock


location time histories in Fig. 35, where the shock location is phase-
locked to the aerofoil motion. Additionally, analysis of the time-
resolved velocity fields and phase-averaged Reynolds stress distribu-
tions indicates the aerofoil motion does not significantly influence the
underlying buffet mechanism, which remains consistent with the authors
earlier findings [113,114].
Raveh & Dowell [133] extended the work on shock buffet lock-in to
spring-suspended aeroelastic systems. In a pitching system, lock-in did
not occur when the structural eigenfrequency was below the buffet. At
these conditions, the moment-coefficient response remained harmonic,
with frequency content concentrated at the rigid aerofoil buffet fre-
quency. The shock oscillation acts as an external forcing to the system,
driving the pitch motion at the buffet frequency. For pitch natural fre-
quencies above the buffet, lock-in occurs following a long transient
beating. Both the aerodynamic and structural responses exhibit only the
pitch natural frequency and converge to large amplitude LCO (Fig. 36(a)
with ζ ¼ 0). Such findings provide further support for the transonic LCO
mechanism described by Iovnovich & Raveh [130], where a phase lead in
the pitching moment for pitch natural frequencies above buffet drives the
instability.
The influence of structural damping on the aeroelastic response of the
pitching aerofoil was also investigated by Raveh & Dowell [133]. As
shown in Fig. 36(a), the addition of a small amount of structural damping
(ζ ¼ 0:5%) is sufficient to quench the lock-in phenomenon for the
particular parameter set considered. A coupled pitch-heave aerofoil has
Fig. 30. Displacement magnitude and phase with respect to excitation of the estimated also been considered by the authors, with structural natural frequencies
shock foot (squares) and the separation point (circles) for three different excitation ki- fα ¼ 1:1fsb and fh ¼ 0:9fsb . Synchronisation of the structural and aero-
nematics at α ¼ 4 (Nitzsche [9]). dynamic responses to the pitch mode is observed, with large amplitude
LCO in pitch as evident in Fig. 36(b). A linear flutter analysis of this
Although the series of studies presented by Raveh and her co-authors coupled system indicates the instability is not the result of flutter. The
are purely numerical, the lock-in mechanism has been evidenced analysis is repeated with the structural natural frequencies reversed (fα ¼
experimentally by Hartmann et al. [131]. Following from earlier exper- 0:9fsb and fh ¼ 1:1fsb ). The results show low amplitude oscillations
iments of the DRA 2303 aerofoil [113,114], the authors investigated exclusively at the buffet frequency, as expected from earlier studies
forced and flow-induced oscillations of the section using TR-SPIV and [125,128,130] where the moment and lift responses exhibit phase lags to
unsteady pressure transducers. In the elastically suspended configura- pitch and heave excitations, respectively. Iovnovich et al. [134] have
tion, coupled pitch and heave motions enable an energy exchange from further suggested that the lock-in phenomenon may be responsible for in-
the fluid to the structure, as dictated by the mean work coefficient of flight LCO encountered by the F-16.
Dietz et al. [132]. For the particular parameter set investigated, such an Recent literature in the field has continued the exploration of aero-
exchange attenuates the amplitude of shock oscillation. Forced pitch/ elastic systems in the presence of shock buffet, concentrating on classi-
heave motion of the aerofoil revealed synchronisation of the shock with fying the influence of various structural parameters on the lock-in
the driving frequency across the various amplitudes and frequencies of phenomenon. In particular, Quan et al. [135] investigated the influence

22
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

both the pitch and lift responses, a consequence of the shock buffet fre-
quency synchronising with the pitch mode.
The sensitivity of the aeroelastic response of a pitching aerofoil to
changes in mass ratio and structural damping observed by Giannelis et al.
[136] also support the findings of Quan et al. [135]. As shown in
Fig. 38(a), the mass ratio does not influence the onset frequency of lock-
in, with the aerodynamic and structural modes synchronising at fα =fSB0 ≈1
in all cases. The results also indicate that sectional mass has little influ-
ence on the LCO amplitude; however, an increase in the mass ratio yields
a narrowing of the lock-in region and reduces the rate of convergence to a
fully developed limit cycle. From Fig. 38(b), the addition of structural
damping to the aeroelastic system imparts greater influence to the
emergence and nature of shock buffet lock-in. With a small degree of
damping, both the LCO amplitude and the extent of the lock-in region
decrease appreciably.
Carrese et al. [73] also investigated the aeroelastic behaviour of the
OAT15A aerofoil, assessing the influence of flight condition, structural
rigidity and natural frequency on the dynamic response of a pitching
aerofoil system. At the nominal flight condition, the findings are in
accordance with previous studies [133,135], whereby lock-in ensues
when the structural natural frequency exceeds that of the buffet. How-
ever, a dependency on the structural rigidity (as dictated by the torsional
stiffness) and the flow condition is also shown. The static trim position of
the aerofoil is inherently linked to both these factors, and in particular, a
lower pitch frequency (corresponding to a lower torsional stiffness) may
yield a static trim position that lies outside the buffet envelope for the
aerofoil. While this may produce a steady aeroelastic response in a
flowfield that would, for a rigid aerofoil, exhibit shock oscillation, the
converse may also apply. Flight conditions that yield a steady aero-
dynamic response for a rigid aerofoil may be perturbed to cross the buffet
boundary when elastic effects are considered. Giannelis et al. [138]
identified similar behaviour, where a coupled pitch-heave OAT15A
aerofoil in pre-buffet conditions is excited by a regulation gust load,
resulting in the system synchronising with the pitch mode and devel-
oping large amplitude LCO.
A series of recent publications by researchers at the Northwestern
Polytechnical University China have also drawn links between the tran-
sonic buffet instability and other aeroelastic phenomena, including single
degree-of-freedom (SDOF) nonlinear flutter [139–141], hereafter deno-
ted SDOF flutter. Gao et al. [142] studied the relationship between SDOF
flutter and buffet on the NACA 0012 aerofoil, free to oscillate in pitch,
through URANS simulation. For structural natural frequencies in the vi-
Fig. 32. Lift-curve slopes vs frequency ratio; α ¼ 1:5 , α ¼ 6 , M ¼ 0:72, Re ¼ 10  106
(Raveh [125]).
cinity of the buffet, and angles of attack up to 1 below onset, LCO can
occur in the pitching system even for eigen frequencies below the buffet,
as shown in Fig. 39. This occurrence is tied to the lock-in phenomenon of
of structural natural frequency, mass ratio and structural damping Raveh [125,133] as the coupled response frequency is at the structural
through URANS simulations for an elastically suspended NACA 0012 in natural frequency if the wind-off pitch frequency is above the buffet. No
pitch. With varying pitch frequency, lock-in first occurs when the struc- qualitative change in the nature of the LCO responses is observed as the
tural and buffet frequencies are approximately equal, and seen to persist angle of attack varies from pre- to post-buffet onset, suggesting a similar
until the pitch frequency is approximately double the buffet. This lock-in mechanism is responsible for SDOF flutter and buffeting.
region contracts with both increasing mass ratio and structural damping. Gao et al. [143] provide further insight into the correlation between
While the upper threshold of the region decreases with an increase in the SDOF flutter and buffet on the NACA 0012 aerofoil, constructing an ARX
structural parameters, the lower bound is insensitive to these changes, reduced-order model for the aerodynamics from URANS simulation and
remaining at approximately the buffet frequency in all cases. analysing the resultant eigenvalue problem for the coupled aeroelastic
In an extension of the earlier work by Giannelis & Vio [74], similar system. Analogous to classical bending-torsion flutter, SDOF flutter is
findings to Quan et al. [135] have been presented by Giannelis et al. found to be the result of modal coupling. However, rather than two or
[136] for the supercritical OAT15A aerofoil. Pitching simulations using more structural modes coalescing, the instability stems from the coupling
the commercial finite volume code ANSYS Fluent [137] indicate the of a structural and aerodynamic mode. For this instability to arise, two
nature of the aeroelastic response of the system is sensitive to the ratio of requisite criteria must be adhered to. Firstly, the fluid must exhibit suf-
structural and aerodynamic frequencies. With reference to Fig. 37, four ficiently low damping i.e. the static aerofoil is at an angle of attack near to
frequency regions of qualitatively distinct behaviour are identified. For buffet onset. Secondly, the instability only occurs for structural natural
fα0 =fSB0 < 0:6, highly nonlinear pitch and lift responses develop, with frequencies close to the buffet, in accordance with the authors earlier
comparatively large amplitude oscillations. In the regions bound by findings [142]. Additionally, two distinct response characteristics are
0:6 < fα0 =fSB0 < 1 and fα0 =fSB0 > 1:8, the pitch response behaves as a single identified for a system experiencing SDOF flutter, dependent on the
degree-of-freedom oscillator, excited by the unsteady shock oscillations. structural natural frequency and mass ratio. At higher structural fre-
Within the frequency band 0:6 < fα0 =fSB0 < 1 resonance is observed in quencies and mass ratios, the instability is governed by the structural

23
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 33. Mach contours along a buffet cycle at α ¼ 6 , M ¼ 0:72, Re ¼ 107 . (a) Beginning of buffet cycle; (b) 27% of buffet cycle; (c) 41% of buffet cycle; (d) 55% of buffet cycle; (e) 61% of
buffet cycle; (f) 68% of buffet cycle; (g) 75% of buffet cycle; (h) 82% of buffet cycle; (i) 90% of buffet cycle (Raveh & Dowell [128]).

Fig. 34. Mach contours along a cycle of heave aerofoil motion of fh =fb ¼ 1:55, α ¼ 6 , M ¼ 0:72, Rec ¼ 107 . (a) Beginning of heave cycle; (b) 21% of heave cycle; (c) 43% of heave cycle;
(d) 64% of heave cycle; (e) 85% of heave cycle; (f) 106% of heave cycle (Raveh & Dowell [128]).

mode, and the flutter frequency locks-in to the structural frequency. At The interactions between classical bending-torsion flutter and tran-
lower structural frequencies and mass ratios, instability in the fluid mode sonic buffet have also been investigated by Zhang et al. [144]. The au-
dominates and the coupled frequency is the characteristic fluid frequency thors considered the Benchmark Active Controls Technology (BACT)
(the buffet frequency at post-buffet conditions). model [145], a rectangular pitch-heave system with the NACA 0012

24
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 35. Shock location and aerofoil pitch at ωf ¼ 0:6942, hf =c ¼ 0:00175, αf ¼ 0:3 and xf =c ¼  0:0515 (Hartmann et al. [131]).

Fig. 37. Effect of frequency ratio on pitch response amplitude and frequency (M ¼ 0:73,
α ¼ 3:5 , Re≈3  106 , μ ¼ 50, ζ ¼ 0%) (Giannelis et al. [136]).

aerofoil section. URANS simulations are performed at M ¼ 0:71, Re ¼


3  106 , α ¼ 0 and a reduced velocity of V  ¼ 0:65, 10% higher than the
reduced flutter velocity. At these conditions, buffet onset occurs at α ¼
4:8 . The resulting response is a form of nodal-shaped oscillations of
Fig. 36. Time histories of aeroelastic response in the present of buffet (Raveh & Dow-
ell [133]).
alternating diverging and damped behaviour, as shown in Fig. 40. The
authors note that as the system is in a supercritical state, the response is
not likely due to a beating superposition of two modes. Rather, the

25
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 39. Amplitudes of responses against the structure reduced frequency kα . Buffet
reduced frequency ksb ¼ 0:18 (Gao et al. [142]).

Fig. 40. Nodal-shaped oscillations of the structural displacements (Zhang et al. [144]).
Fig. 38. Effect of mass ratio and structural damping on pitch response amplitude (Gian-
nelis et al. [136]). with variations in the structural natural frequency reveal the presence of
mode veering [147,148]. At pitch frequencies below and well in excess of
oscillations result from the interaction between the flutter and buffet the buffet frequency, the aeroelastic system undergoes forced vibration at
modes. As shown in Fig. 41, when the pitch angle exceeds the buffet onset the buffet frequency. For structural frequencies approximately equal to
angle, the lift response begins to exhibit higher frequency components, the buffet frequency, an exchange between the structural and fluid modes
the most dominant of which is the buffet frequency. The change in occurs, with the aeroelastic response following neither the buffet nor
aerodynamic loading causes the structural modes to dissociate, migrating pitch frequencies. Within the lock-in region, coupling between aero-
from the flutter frequencies to their respective natural frequencies. This is dynamic and structural modes results in instability in the pitch degree of
accompanied by an abrupt phase reversal between the two structural freedom and the response tracks the structural natural frequency. In
modes, inhibiting the energy transfer between the fluid and structure and accordance with their previous work, Gao et al. [146] assert that lock-in
producing the damped oscillations evident in Fig. 41 at approximately is a consequence of a form of SDOF flutter, where the instability results
t ¼ 3 s. from the coupling of the structural system with an unstable fluid mode.
The emergence of nodal-shaped oscillations in the work of Zhang
et al. [144] prompted further analysis of the mechanism governing buffet 6. Three-dimensional transonic shock buffet
lock-in by Gao et al. [146]. Linear stability analysis of the coupled NACA
0012 aeroelastic system is performed, employing the reduced-order 6.1. Experimental investigations
aerodynamic model detailed by Gao et al. [142]. The authors note that
the asymmetry of the lock-in region about the buffet frequency is not The investigations of transonic shock buffet on aerofoils outlined in
representative of a pure resonance phenomenon, as was observed in the preceding sections surmise that the governing physics of the phe-
vortex-induced vibrations of an oscillating cylinder for wind-off struc- nomenon is essentially two-dimensional. As discussed by Jacquin et al.
tural frequencies both above and below the characteristic flow frequency. [20], even in the case of a constant cross-section, rectangular wing sec-
Rather, analysis of the structural and least stable fluid mode eigenvalues tion, the dimensionality of the buffet instability is questionable. Early

26
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

where incipient and massively separated flow is present. Steady and


unsteady pressure transducers are positioned at various spanwise sta-
tions, with static pressure orifices along a chordwise strip at 28% span.
Additionally, fast-response anodic aluminium pressure-sensitive paint
(AA-PSP) is applied to the wing surface for temporal resolution of the
surface pressure distributions. The aeroelastic response of the wing is
recorded through videogrammetric deformation measurement at two
flow conditions; condition 1 (½α; M ¼ ½0 ; 0:86) in which a small degree
of trailing edge separation is present and condition 2 (½α; M ¼ ½0 ; 0:92)
where massive shock-induced separation is observed (consistent with
Pearcey Type-B3 separation [8]).
The dynamics of the unsteady flowfield, and consequently, the
structural response, at the two conditions considered each exhibit unique
characteristics. In the weak shock-wave/boundary layer interaction
condition 1, spectral analysis of the pressure fluctuations reveals several
harmonic peaks. While some of these peaks represent harmonics of the
test section, a characteristic shock oscillation frequency is present. The
authors attribute the high level of unsteadiness at this condition to an
aeroacoustic feedback loop between the shock and trailing edge pressure
perturbations. Time sequences of surface pressure distributions, such as
Fig. 41. Pitching displacements and lift coefficient in one cycle of nodal-shaped oscilla- those of Fig. 49, show shock motion concentrated at the wingtip, corre-
tion (Zhang et al. [144]). lating well with numerical studies [152–155]. The shock oscillations also
exhibit a highly aperiodic character, with broadband structural oscilla-
tions appearing at outboard stations. The strong shock-wave/boundary
experiments of buffeting flows on three-dimensional wings, although layer interactions of condition 2 produce a relatively steady separation
finding some degree of commonality in the underlying fluid-dynamic line, owing to the inertia of the shock-induced separation. While this
phenomena, highlighted distinct behaviours that are unique from the reduces the levels of pressure fluctuations, the structural response is
two-dimensional buffet characteristics observed on aerofoils. Roos [149], excited at the dominant aerodynamic frequency in a resonant response
for instance, examined unsteady pressure fluctuations over a high aspect analogous to that described by Nitzsche [9].
ratio, transport-type swept wing configuration in the NASA-ARC 14-foot Lawson et al. [156] have detailed an experimental programme con-
transonic wind tunnel. From the experiments, chordwise pressure ducted under the Buffet Control of Transonic Wings (BUCOLIC) collab-
perturbation convection velocities were found to be consistent with orative project between the Aircraft Research Association and the
earlier studies on two-dimensional aerofoils [108]. However, deep University of Liverpool. The experiments investigated transonic flow
within the buffet region, large-scale unsteadiness is found to be most over the RBC12 half wing-body model, a variant of the B60 model used to
severe at the wingtip. This is contrary to the predominantly constant study nacelle installation effects [157] and a representative geometry of
spanwise distribution of pressure fluctuations observed in experiments modern civil transport aircraft. The wing is equipped with a variety of
over full-span aerofoil models. Additionally, the characteristic flow fre- measurement apparatus, including steady and unsteady pressure trans-
quencies were found to be approximately an order of magnitude higher ducers, dynamic pressure-sensitive paint (DPSP), accelerometers and
than those of a two-dimensional aerofoil, with pressure spectra exhibit- wing root strain gauges, which are employed to assess the efficacy of a
ing broadband bumps for Strouhal numbers in the vicinity number of commonly used buffet indicators for onset prediction.
of 0:2 < St < 0:4. The buffet indicators considered are derived from recommendations
These broadband spectral characteristics have also been noted by put forth by the ESDU [158], and include wing root strain divergence,
Benoit & Legrain [150] in transonic experiments of shock buffet over a RMS wingtip acceleration, lift-curve slope break, pitching moment break,
transport-type wing in the ONERA Chalais Meudon S3Ch wind tunnel. axial force divergence and trailing edge pressure divergence. The onset
The authors further investigated the buffeting flowfield of a constant predictions of each of the indicators is summarised in Fig. 50 across a
RA16SC1 cross-section wing of moderate aspect ratio with zero and range of Mach numbers. As shown, all indicators excluding axial force
moderate sweep. In the rectangular wing case, shock oscillations were divergence and pitching moment break give consistent results
seen to originate at the root and extended outboard with increases in throughout the Mach number range considered. A further comparison
incidence. Although three-dimensional effects contributed to less between the strain gauge divergence and DPSP onset predictions is given
organised flow dynamics, narrowband pressure spectra were obtained in Fig. 51. The onset of buffeting through strain gauges is 0:5 below the
with peaks corresponding to the frequencies observed in experiments of a onset of large-scale flow unsteadiness determined from the DPSP mea-
two-dimensional RA16SC1 aerofoil. Overall, the nature of the three- surements. This is rectified by considering the frequency range for which
dimensional buffet instability for the unswept wing was found to corre- pressure oscillations develop. While broadband fluctuations of the
late well with the experiments on two-dimensional sections. Conversely, flowfield develop at α ¼ 3 and coincide with onset through DPSP
results of the swept wing configuration are indicative of distinct flow measurements in Fig. 51, pressure oscillations in the frequency band of
phenomena. As for transport-type wings, shock unsteadiness first 75–150 Hz originate at α ¼ 2:7 . The strain gauge onset is thus a result of
emerges at the wingtip and progresses inboard with incidence. At excitations in this lower frequency band, where the structure is more
developed buffet conditions, pressure fluctuations are again of highest responsive to external forcing. The authors suggest that at onset, the
intensity at the wingtip. These fluctuations are broadband in nature, and structure is more receptive to pressure fluctuations due to incipient
present a qualitative distinction to the organised and periodic buffet separation, rather than large-scale shock oscillations.
cycles observed on aerofoils. Further analysis of the unsteady pressure transducer and DPSP data
More recent experiments of three-dimensional shock buffet have acquired by Lawson et al. [156] has been undertaken by Masini et al.
verified the early findings of Roos [149] and Benoit & Legrain [150]. [159]. Proper orthogonal decomposition of the DPSP snapshots has been
Steimle et al. [151], for instance, studied unsteady transonic flow over a employed to identify dominant features of the buffeting flow. In Fig. 52,
flexible transport-type swept wing with BAC 3-11/RES/30/21 sectional the first four modes of a clean wing configuration are shown, with the
geometry, assessing the fluid-structure coupling mechanism at conditions highest energy mode associated with the structural response of the wing.

27
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Higher frequency modes are also identified, which correspond to shock- 12 unsteady pressure transducers along the suction surface of the wing at
induced unsteadiness concentrated at various spanwise bands. In 50% and 60% span. Steady pressure readings were acquired at an addi-
particular, modes 2, 4 and 5 exhibit high spatial amplitudes at inboard, tional six spanwise stations and aerodynamic forces measured by the
central and outboard spanwise stations, respectively. A dependence of tunnel load balance [164]. Additionally, aeroelastic deformations of the
the characteristic frequencies of each mode on the local chord is also wing are provided through videogrammetric recording, such that wing
identified, with the frequency content of shock unsteadiness migrating to twist and bending along the span is available at each incidence. Trip dots
higher frequencies at outboard stations. Additionally, by reconstructing are installed at 10% chord on the suction surface and data is presented for
the pressure field through combination of the aerodynamic modes, the incidence sweeps at M ¼ 0:85 with Re ¼ 0:949  106 and Re ¼
authors have provided insight into the nature of spanwise shock dy- 1:515  106 . As the distribution of sensors is insufficient to completely
namics. At angles of attack in the vicinity of onset, inboard propagating capture shock motion at the lower Reynolds number condition, the au-
pressure perturbations are observed from approximately 87% span, thors direct their discussion to the results recorded at Re ¼ 1:515  106 .
running along the shock from tip to root. These spanwise perturbations The qualitative flow features identified by Koike et al. [163] are again
are, however, dependent on incidence. At developed buffet conditions an consistent with earlier experiments of three-dimensional shock buffet. In
additional outboard propagating perturbation also appears, which is particular, the pressure spectra reveal that with increasing incidence,
associated with the higher-frequency bumps in the pressure spectra. three distinct characteristic responses emerge:
Dandois [160] has presented the results of two wind tunnel tests
conducted at ONERA for the investigation of closed-loop buffet control 1. At low incidence (α < 3 ), the shock is essentially stationary and low
under the Aerodynamic Validation of Emission Reducing Technology intensity pressure fluctuations are evident, with resolved spectral
(AVERT) project. The AVERT model is a swept half wing-body configu- peaks corresponding to the tunnel natural frequencies.
ration of OAT15A aerofoil section and with sufficient rigidity to produce 2. At moderate incidence (3 < α < 5:5 ), small amplitude shock oscil-
negligible oscillations in the bending and torsion modes. Test are con- lations are present with a spectral bump at St≈0:3. This spectral bump
ducted for Mach numbers between 0.78 and 0.86, with Reynolds number corroborates the findings of Roos [149] and Dandois [160].
(based on mean aerodynamic chord) ranging between 2:83  106 and 3. At high incidence (α > 5:5 ), large amplitude, aperiodic shock oscil-
8:49  106 . The model was instrumented with 86 static pressure tappings lations are observed, which exhibit low-frequency, broadband
at four spanwise stations, 57 unsteady pressure transducers at seven spectra. Pressure spectra of this form are in accordance with the ex-
spanwise sections and three stations equipped with two accelerometers periments of Benoit & Legrain [150].
each. A sublimating oil film was also applied to the upper wing surface to
track separation characteristics. Cross-correlation and analysis of coherence between unsteady pres-
The investigations explicitly considered the effects of both Mach sure measurements at the two spanwise stations by the authors has
number and Reynolds number on buffet onset. The effect of an increase in provided further insight into the nature of spanwise pressure fluctuations
Reynolds number is to delay buffet marginally throughout the regime on three-dimensional wings. For both moderate and high incidence
considered in the experiments. This delay is independent of incidence conditions, spanwise undulations originated near the wing root and
and is equal to 0:1 at Re ¼ 4:72  106 and 0:4 at Re ¼ 8:49  106 , as propagated outboard towards the tip. The convection speeds of these
compared to the baseline results at Re ¼ 2:83  106 . Mach number ef- spanwise perturbations decreases with an increase in angle of attack.
fects on onset are more pronounced. A higher Mach number is accom- These findings have been complemented by analysis of fast-response PSP
panied by a downstream shock displacement, with onset incidence data for buffeting flow over the CRM detailed by Sugioka et al. [165].
decreasing with an increase in Mach number.
Consistent with the findings of Roos [149] and Benoit & Legrain
[150], the experiments of Dandois [160] have questioned whether buffet 6.2. Numerical investigations
in a three-dimensional sense can be explained by the two-dimensional
models described in Section 2. From the PSDs of pressure fluctuations Exploration of the three-dimensional buffet mechanism through wind
shown in Fig. 53, the buffet instability appears as a broadband bump at tunnel experiments have been supplemented by a number of extensive
frequencies between 0:2  St  0:6. In the experiments of an equivalent computational investigations of the phenomenon. These studies have
two-dimensional wing section performed by Jacquin et al. [20], the further indicated that three-dimensional effects may significantly alter
instability appears as a well-resolved peak with St ¼ 0:06. A physical the nature of shock oscillations, which are particularly sensitive to
explanation for this disparity is offered by the spanwise propagation of planform geometry [20,161]. Owing to advances in computer processing
so-called buffet cells described by Iovnovich & Raveh [161] (see Section power, numerical simulation through URANS or scale-resolving methods
6.2). The PSDs in Fig. 53 further highlight the presence of a secondary has become an indispensable tool in probing the three-dimensional na-
Kelvin-Helmholtz type instability, manifesting as a broadband bump ture of the buffet phenomenon.
between 1  St  4. This secondary instability has also been noted in the Brunet & Deck [152] have provided one of the earliest studies on
global stability analysis performed by Sartor et al. [29], and represents a three-dimensional shock buffet simulation. The Zonal-DES method of
shear layer interaction downstream of the shock. Deck [18] described in Section 3 is employed in ONERA's elsA code [57]
For investigation of the three-dimensional buffet mechanism, the to simulate shock buffet of the CAT3D half wing-body model studied
research community has shown significant interest in the NASA Common experimentally by Caruana et al. [166]. Computations are performed at
Research Model (CRM) geometry [162]. The CRM is a generic transport- M ¼ 0:82, Re ¼ 2:8  106 and α ¼ 4:2 , a condition for which established
type aircraft developed by NASA for the validation of numerical routines, buffet was observed in the experiments. DES computations are limited to
and has been the subject of extensive study in the American Institute of regions of separated flow, encompassing the entire wing-body wake and
Aeronautics and Astronautics Drag Prediction Workshops. Buffet exper- the outboard portion of the upper wing surface. The simulations suc-
iments have been explicitly conducted on an 80% scaled variant of the cessfully capture the massively separated and unstable regions on the
CRM by Koike et al. [163] in the 2 m  2 m transonic wind tunnel of the outboard 50% of the wing. As seen in Fig. 42(a), Kelvin-Helmholtz type
Japanese Aerospace Exploration Agency (JAXA). In a similar light to the instabilities become prominent downstream of the shock, with eddy roll-
experiments conducted by ONERA on the OAT15A aerofoil, the wind up observed in the separated region and a distortion of the shock towards
tunnel tests performed by JAXA look to establish a comprehensive the tip. Fig. 42(b) further highlights these effects, where the maximum
database for the validation of three-dimensional buffet simulations. pressure fluctuations are due to shock motion. The curved shock pattern
In the experiments of Koike et al. [163], the model was equipped with is a distortion of the steady flow resulting from the massively separated
zone that produces an upstream shock displacement.

28
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 42. Visualisation of pressure waves (Brunet [152]).

Results of the DES calculations correlate well with the experiment in


terms of mean and RMS pressure distributions along the span. Compar-
isons between RANS simulations including a variety of turbulence
models and grid densities showed that in all instances the RANS com-
putations underestimated the levels of separation. The authors suggest
that accurately resolving turbulent scales may be a necessity to model
correctly the three-dimensional buffet instability. Analysis of the
Fig. 43. Visualisation of pressure waves (Brunet [152]).
instantaneous dilation field ∇ðρuÞ in Fig. 43 reveals apparent pressure
wave propagation from the trailing edge. This radiation process is
inherently three-dimensional; large-scale structures impinge on the SST model predicts the largest extent of separation. Although the various
trailing edge and generate acoustic waves which propagate upstream closures result in different estimates for the critical angle at which large-
(Fig. 43(a)) and along the span (Fig. 43(b)), a dispersion which affects the scale separation (and implicitly, buffet onset) occurs, each produces
dynamics of the separation line. similar flow features. Nonetheless, comparisons between steady and
Sartor & Timme [153] have performed RANS and URANS computa- unsteady simulations show that both the extension of shock-induced
tions using a number of turbulent closures with the DLR-TAU code [61], separation to the trailing edge and the appearance of a kink in the lift
confirming the shock motions on the outboard portions of the wing as a curve slope, two commonly applied indicators of buffet onset, are unre-
characteristic of three-dimensional shock buffet. The RBC12 half wing- liable measures for the emergence of large-scale unsteadiness.
body model is investigated, corresponding to the geometry studied by Unsteady simulations found that both the Spalart-Allmaras and
Lawson et al. [156]. A number of turbulence models were considered, Menter's SST models provide similar predictions for onset (α≈3:1 ),
including the negative Spalart-Allmaras [32], Menter's k ω SST [52], a although the severity of buffet is highly dependent on the particular
k ω LEA [63] and an explicit algebraic Reynolds stress model in the model chosen. The LEA and RQEVM models produced similar results to
form of a Realisable Quadratic Eddy Viscosity Model (RQEVM) [63]. At one another, but predict onset at a higher incidence relative to the pre-
small incidence, each of the models produced consistent predictions of vious closures (α≈3:8 ). PSDs of the lift signal also showed that onset is
shock-induced separation. At higher angles of attack, the Spalart- dominated by narrowband frequency content consistent with periodic
Allmaras and RQEVM models yield similar lift predictions, while the motion. At higher angles of incidence, the instability adopts a broadband

29
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

character characterised by aperiodic motions. Whereas Brunet & Deck


[152] suggested that scale-resolving simulations may be necessary to
capture three-dimensional buffet, Sartor & Timme [153] have shown that
the URANS approach is capable of reproducing the dominant flow fea-
tures. With increasing incidence, the unsteady shock motions first
become apparent at the wingtip and grow in spanwise extent towards the
fuselage. The increase in incidence is also accompanied by the develop-
ment of broadband frequency content, with the dominant frequency
reducing as the shock unsteadiness spreads to spanwise stations of larger
chord. This highlights a fundamental distinction between two- and three-
dimensional buffet characteristics, where at developed buffet conditions
for a three-dimensional geometry, shock oscillations typically exhibit
aperiodic motions.
Subsequent studies by Sartor & Timme [154,155] have compared the
prior URANS results to DDES computations. Both methods show excel-
lent agreement regarding shock position inboard of the crank, however,
the outboard mean shock foot trace of the DDES results exhibit a
straighter character, more consistent with the experimental results of
Lawson et al. [156]. Standard deviations of the surface pressure distri-
butions derived from DDES computations are also in better agreement
with the experiments, supporting Brunet & Deck's [152] notion that
scale-resolving simulations may be necessary to obtain a comprehensive
description of the fundamental flow physics. Nonetheless, both methods
are able to reproduce self-sustained shock oscillations outboard of the
crank. From Fig. 44, turbulent structures are seen extending from the
shock foot in both cases, with the separated zones shifting the shock foot
toward the leading edge. Due to the resolution of turbulent eddies, the
DDES computations show a rich turbulent spectrum, which also mani-
fests in the time histories of the aerodynamic coefficients. Consistent
with the experiments of Koike et al. [163], the separated regions migrate
from inboard to outboard sections, producing regular yet aperiodic os-
cillations of the integral forces. PSDs of the force coefficients further
demonstrate this aperiodic behaviour, with low-frequency broadband
peaks between 150 and 300 Hz in both approaches. In analysing
instantaneous velocity divergence fields, the authors came to similar
conclusions as Brunet & Deck [152], finding that acoustic wave propa-
gation is an inherently three-dimensional process where both streamwise
and spanwise radiation of disturbances influence shock and separa-
tion dynamics.
Extending their earlier work, Sartor & Timme [167] studied the effect
of Mach number on the emergence of unsteady phenomenon on the
RBC12 geometry. RANS simulations found that at smaller Mach numbers,
the appearance of a kink in the drag polar provided a fair indicator of the
onset of shock buffet. This break in the polar is, however, an unreliable
metric, a consequence of unconverged steady simulations that does not
appear at higher Mach numbers. Unsteady simulations across a range of
conditions showed that onset incidence decreases with Mach number.
Additionally, higher Mach number simulations exhibit lower magnitude
pressure fluctuations, with unsteadiness concentrated in the separated
flow regions and the standard deviation of lift oscillations plateauing
soon after onset. Conversely, although the extent of separated flow at
lower Mach numbers is reduced, unsteadiness develops across the en-
tirety of the suction surface and is accompanied by large amplitude Fig. 44. Instantaneous values of eddy-viscosity ratio at α ¼ 3:8 , M ¼ 0:80 and Re ¼
pressure fluctuations. 3:75  106 (Sartor & Timme [154]).
With confidence in the efficacy of URANS methods to produce the
main flow features associated with three-dimensional shock buffet, Iov- buffet amplitudes and frequencies are consistent between two- and three-
novich & Raveh [161] conducted a numerical parametric study to assess dimensional simulations, and are in fair agreement with the predomi-
the effects of sweep angle and span on the buffet instability mechanism. nantly two-dimensional mechanism for infinite-straight wings described
Using the finite difference RANS solver EZNSS [31], three baseline by Jacquin et al. [20].
configurations were investigated: infinite-straight, infinite-swept and Analysis of the buffet behaviour of infinitely-swept wings at variable
finite-swept wing models, each with a constant RA16SC1 section. For the sweep angles revealed ranges of sweep for which qualitatively distinct
infinite-straight wing model, nonuniform lateral structures develop flow mechanisms are observed. At low sweep angles (Λ < 20 ), the buffet
during upstream shock excursions and the shock travels farther forward mechanism is similar to the infinite-straight and two-dimensional cases,
than in the two-dimensional case due to the increased separation asso- with outboard shock oscillations primarily along the chord and minimal
ciated with three-dimensional effects. Nonetheless, excluding the local- lateral pressure disturbance propagation. The surface pressure RMS
ised three-dimensional character during parts of the oscillation cycle, the contours in Fig. 45 show that with increasing sweep in this regime, the

30
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 45. Surface pressure RMS levels for various sweep angles: infinite-swept configuration at nominal buffet conditions (Iovnovich & Raveh [161]).

shock travel becomes increasingly nonuniform along the span. For two time instances. The authors describe the physical mechanism gov-
moderate sweep angles (20  Λ < 40 ), a lateral flow mode begins to erning the propagation of these buffet cells as follows:
dominate, which eliminates the two-dimensional buffet instability
through pressure perturbations propagating outboard from the wing 1. Pressure fluctuations are generated at the aft λ shock of the wing root
root. The unsteady shock motions observed at moderate sweep are and emanate outboard.
distinct from the two-dimensional buffet mechanism and are charac- 2. As a positive pressure fluctuation propagates to the spanwise station
terised by the periodic spanwise convection of so-called buffet cells, as of the λ shock tip, unstable shock-wave/boundary layer interaction
evident in the surface pressure fluctuation maps of Fig. 46 for Λ ¼ 30 at ensues due to an increase in local shock strength, resulting in an

Fig. 46. Surface pressure coefficient and fluctuation maps at two time snapshots during shock-buffet oscillations: Λ ¼ 30 infinite-swept configuration and nominal buffet conditions
(Iovnovich & Raveh [161]).

31
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

upstream displacement of the shock and an increase in shock-induced frequency-domain analysis has been performed on the RBC12 half wing-
separation. body configuration at conditions approaching buffet onset using the DLR-
3. The corresponding negative pressure fluctuation also propagates from TAU code [61]. The authors note that the high dimensionality of the fluid
the wing root along the aft λ shock, but does not yield an increase in Jacobian prohibits the use of direct methods for eigenvalue calculations
shock strength. As such, as the perturbation passes the λ shock tip, it due to inordinate memory requirements. Additionally, the stiffness of the
propagates outboard along a constant (aft) chordwise position, with linearised aerodynamics near the onset of instability is problematic;
flow remaining attached aft of the shock. conventional iterative methods, such as the incomplete lower-upper
4. With both positive and negative pressure fluctuations emanating (ILU) preconditioned generalised minimal residual (GMRES) approach
periodically from the wing root, outboard propagating buffet cells [170], are particularly susceptible to stall. These issues are circumvented
result, which at an individual spanwise stations exhibit the chordwise by use of an advanced iterative solution scheme, the generalised conju-
shock motion characteristic of two-dimensional buffet, but which gate residual solver with deflated restarting (GCRO-DR). A complete
stem from an inherently three-dimensional phenomenon. description of the theory and implementation of the method is provided
by Xu et al. [171].
This spanwise propagation of buffet cells appears to be tied to a Timme & Thormann [169] performed frequency sweeps at four an-
higher-frequency (relative to two-dimensional buffet) aerodynamic gles of attack below buffet onset (α ¼ 3:03 ), using a synthetic torsional
mode, as indicated by the incremental mode decomposition conducted mode for excitation. As shown in the lift coefficient frequency response of
by Ohmichi et al. [168] on the CRM configuration. Fig. 48, as the incidence approaches onset, a distinct resonant peak
In addition to a change in the nature of unsteadiness, Iovnovich & emerges at St ¼ 0:11. Within this low-frequency range, the aerodynamic
Raveh [161] found that an increase in sweep is also accompanied by a response leads the structural excitation, a potential indicator of a desta-
reduction in buffet amplitude (as shown in the RMS contours of Fig. 45) bilising aerodynamic mode that was also observed in two-dimensional
and an increase in frequency. A further increase in sweep (Λ  40 ) stability analysis of aerofoils. Contrary to the two-dimensional case,
yields a stationary shock position due to stall. From these results, the however, interesting response characteristics emerge for Strouhal
authors relate the effects of sweep angle to incidence. An increase in numbers between 0.3 and 0.7. Marginal increments in incidence result in
sweep, as with an increase in incidence, results in an increase in buffet the appearance and amplification of secondary peaks, which coincide
frequency, up to the point at which the wing stalls and shock oscilla- with the typical frequency band for which broadband spectra have been
tion ceases. reported in buffeting flows on three-dimensional wings and which may
Regarding the finite-swept wing configurations, tip-vortex in- reflect the presence of additional unstable aerodynamic modes.
teractions were found to contribute to the buffet mechanism. The mean The complex-valued unsteady pressure distribution has also been
upper surface pressure coefficients for variable span length at a nominal examined, in particular, for excitation frequencies corresponding to the
sweep of Λ ¼ 20 are shown in Fig. 47. For b=c ¼ 2, pressure fluctuations low-frequency peak (St ¼ 0:11) and the dominant secondary peak (St ¼
were limited to the tip, where tip-vortices dominate and shock buffet 0:51). At St ¼ 0:11, unsteadiness is concentrated at the outboard shock
does not develop. At longer span extents, the lateral propagation mech- foot. The pressure fluctuations grow in spanwise extent and intensity
anism found in infinite-swept wings is also observed; however, the with an increase in incidence, consistent with prior experimental and
mechanism is somewhat more complex due to the interaction between computational observations [150,153]. At the higher-frequency excita-
the outboard pressure fluctuations due to shock oscillation and the tion, the unsteady flow exhibits a distinct character. Minimal pressure
wingtip vortices. The main oscillation character along the span is fluctuation was observed at α ¼ 2:9 ; however, upon closer approach to
consistent with the infinite-swept wing results, with irregular shock onset, high intensity unsteadiness is evident at the wingtip shock foot and
motion emerging at outboard stations. downstream of the shock. Relative to the low-frequency excitation, the
While Iovnovich & Raveh [161] have provided a physical description shock centred unsteadiness is concentrated at the wingtip. Considering
of the three-dimensional buffet mechanism, the recent work of Timme & explicitly the eigenspectrum across the various angles of attack, an in-
Thormann [169] has shown that the origins of the instability may lie in crease in incidence is accompanied by the migration of several eigen-
globally unstable aerodynamic modes; an analogue to the two- values in the frequency range 0:3 < St < 0:7 towards the imaginary axis.
dimensional instability detailed by Crouch et al. [21]. Linearised Again, the distinct behaviour in this higher-frequency band suggests a

Fig. 47. Surface pressure coefficient for various span length cases: Λ ¼ 20 finite wing configurations at nominal buffet conditions (Iovnovich & Raveh [161]).

32
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Koike et al. [163]. In a similar light, integration of the Zonal-DES method


of Deck [18] into the JAXA code FaSTAR [177] by Ishida et al. [178] has
also produced favourable correlations to the CRM experimental data,
albeit with shock locations displaced upstream. The CRM configuration
has also been subject to aerodynamic shape optimisation with buffet
constraints by Kenway & Martins [179], in which the optimisation suc-
cessfully drove the buffet boundary of the optimal configuration away
from the design-point.

7. Control of transonic shock oscillations

7.1. Trailing edge deflections

A common feature among the various models governing the transonic


buffet phenomenon presented in Sections 1 and 2 is a coupling between
the large-scale excursions of the shock and pressure fluctuations at the
trailing edge. It follows that control of this instability can be achieved by
modulating the flowfield at either the shock-wave/boundary layer
interaction zone or at the trailing edge. For the latter, the efficacy of
trailing edge flap deflections for the control of shock oscillations have
been investigated in the classical literature [180,181]. While static
trailing edge flap deflections may effectively alter the flow state to a
stable region, they are often accompanied by detrimental aerodynamic
effects, particularly appreciable drag penalties, rendering these methods
undesirable for the control of shock oscillations.
Under the BUFET’N Co [182] project at ONERA, a Trailing Edge
Deflector (TED) was developed to take advantage of the shock stabili-
sation achieved by modulating flow at the trailing edge while minimising
the effects on the global flow characteristics. Despre et al. [124] detail the
design of the TED (as shown in Fig. 54), where a servo-motor is con-
nected to the tab (spanning 1–3% chord) on the pressure surface of
the model.
Various control strategies employing the TED were investigated both
experimentally and computationally using the viscous-inviscid interac-
tion code VIS15 [183,184] on the two-dimensional OAT15A section.
Static TED deflections were shown to reduce the incidence of buffet
onset, however, the lift at onset increased. Open-loop sinusoidal de-
flections at developed buffet conditions with particular phase shifts were
briefly able to stabilise the shock oscillations. It was observed that an
increase in the deflector angle resulted in downstream shock excursions,
and vice versa. Nonetheless, closed-loop control was deemed necessary
for a robust control strategy. Using unsteady pressure signals at the rear
of the shock-wave/boundary layer interaction zone, the control law
involved positive TED deflections during upstream shock excursions and
negative TED deflections during downstream shock travel - actuation that
opposes the natural motion of the shock. For a pressure signal P, the TED
deflection δðtÞ is governed by:
 
Fig. 48. Frequency response of lift coefficient due to flexible torsion mode (Timme & δðtÞ ¼ δ þ A Pðt  τÞ  P (13)
Thormann [169]).

where δ and P are the mean TED deflection and pressure measurement,
different physical mechanism underlying three-dimensional buffet. respectively. The amplitude A and time delay τ were parameters that
The preceding discussion illustrates the fundamental advances in the were varied throughout the investigation. With an optimal set of pa-
understanding of three-dimensional shock buffet attributed to compu- rameters, Despre et al. [124] saw significant reductions in the severity of
tational efforts. However, these investigations are supplemented by an shock buffet through active TED control, a decrease in shock oscillation
array of additional numerical studies, which act to corroborate the un- amplitude of approximately 66% in the experiments.
derlying fluid-dynamic phenomena. Xiong & Liu [172] have employed A subsequent study by Caruana et al. [185] employing the TED for
DES to successfully capture three-dimensional shock buffet on a swept buffet control identified similar static behaviour, and with small modi-
wing with a supercritical aerofoil, providing further evidence to the fications in the control law achieved comparative control effectiveness to
three-dimensional nature of pressure wave propagation. Illi et al. Despre et al. [124]. As shown in Fig. 55(a), static TED deflections delay
[173,174] employed both URANS and DDES computations on the CRM the onset of buffet to higher lift coefficients, however, in the developed
geometry to show that the large vortices shed from the main wing at buffet region this control strategy is ineffective at reducing the magni-
buffet conditions have a significant influence on the flowfield at the tude of pressure fluctuations. The efficacy of closed-loop TED deflections
horizontal tailplane. Ribeiro et al. [175] have detailed modifications to is clearly evident in Fig. 55(b), where the amplitude of shock travel is
the Lattice-Boltzmann based commercial solver PowerFLOW [176], substantially lower with the control law active.
which allow excellent buffet predictions relative to the experiments of Caruana et al. [186] further investigated the effectiveness of TED

33
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 49. Representative pressure distribution time sequence on the upper wing surface measured by AA-PSP, condition 1 flow ½α; M ¼ ½0 ; 0:86, Re ¼ 1:05  106 , Δt ¼ 0:667 ms. (Steimle
et al. [151]).

Fig. 50. Comparison of buffet onset indicators (Reproduced by kind permission of Aircraft
Research Association Ltd Lawson et al. [156]). Fig. 51. Comparison of buffet strain gauge response with DPSP and CFD unsteadiness
(Reproduced by kind permission of Aircraft Research Association Ltd Lawson et al. [156]).

control on a three-dimensional wing-fuselage model experimentally at


ONERA's S2Ma wind tunnel. Static TED deflections provide an increase in Sinusoidal TED deflections in open-loop on a three-dimensional wing
aerodynamic performance analogous to the two-dimensional effects, are qualitatively different from those of the two-dimensional section
delaying onset to higher values of lift. While drag penalties are incurred previously investigated. Caruana et al. [186] found that the presence of
at lower lift coefficients, the authors note that multiple TED actuators three-dimensional separated flow structures must be considered, with
along the span may be employed to impart artificial wing twist, allowing open-loop deflections in many cases exacerbating the shock motions and
optimal configurations to minimise lift-induced drag at various flight increasing levels of structural vibration, particularly in torsion. The
conditions. inherently three-dimensional flowfield and non-uniform spanwise shock

34
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 52. Spatial component of dominant modes related to the structural response and spanwise shock unsteadiness. The spatial amplitudes are coloured from blue to red, representing
opposite signs. The numbers in parentheses represent the POD energy of the mode and the cumulative POD energy from the first mode until the respective mode (Masini et al. [159]). (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

location also poses issues for closed-loop TED control, particularly as the an effective means of buffet alleviation. Numerical simulations of the
principle flow direction does not coincide with the axis of the deflectors NACA 0012 aerofoil at developed buffet conditions and with a trailing
and transducers. By low-pass filtering of the measured pressure signals, edge flap axis of rotation at 80% chord are performed. The lift coefficient
the authors demonstrated computationally an effective means of con- Cl is employed to determine the present state of the system due to the
trolling shock oscillations on the outboard sections of the wing by TED relative smoothness compared to pressure fluctuations or shock location.
deflections through the central span. The resultant shock displacements The flap deflection βðtÞ is governed by the control law:
and commanded TED deflection are shown in Fig. 56. Caruana et al.
[186] do note, however, that such a control strategy may be detrimental βðtÞ ¼ λ½Cl ðt  τÞ  Cl0  (14)
to the flow over the central span and further investigations are required
with more complex control laws. where Cl0 is the balanced lift coefficient, λ is the dimensionless gain and τ
Although the use of trailing edge flap deflections may be accompa- is the time delay in nondimensional time. Exploring the influence of the
nied by various detrimental aerodynamic effects, Gao et al. [187] have various parameters, the optimal balanced lift coefficient is found to be
demonstrated that with a closed-loop control law this form of actuation is that pertaining to the unstable steady solution and is obtained by

35
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 53. PSDs for different chordwise locations at 60% span; M ¼ 0:82, Re ¼ 2:5  106
(Dandois [160]).

Fig. 55. Effects of static and dynamic TED deflections of OAT15A aerodynamic charac-
teristics (Caruana et al. [185]).

the three control parameters, shock oscillations are completely quenched


with reasonable control effort, as shown in Fig. 58.
Although not strictly related to trailing edge deflections, Liu & Yang
[188] have shown, computationally, that modification of the suction
surface flowfield through installation of a wall-normal protrusion aft of
the shock may provide an effective means of buffet load alleviation. The
NASA SC(2)-0714 aerofoil, studied experimentally by Bartels & Edwards
[189], is employed as a baseline from which subsequent simulations
investigating the effect of chordwise location and protrusion height of the
microtab were developed. The installation of such a passive control de-
vice imparts a geometric effect on the flowfield (as discussed in Section
2.3). The microtab alters the dynamics of the main vortices shed from the
shock-wave/boundary layer interaction by introducing a secondary
vortex. The height difference between vortices fore and aft of the
Fig. 54. Trailing edge deflector (TED) (Caruana et al. [185]).
microtab results in a tendency of velocity variations, which (for certain
positions of the protrusion) weaken interactions aft of the shock. The
iteratively varying the flap deflections until Cl0 equals the unstable steady
authors found a microtab positioned at 80% chord with a protrusion
lift coefficient. The authors note that this value differs from the time-
height of 0.75% chord was the most effective configuration for buffet
averaged lift coefficient and that the variation in drag incurred from
load alleviation, reducing buffet amplitudes by 93% while shifting fre-
these control deflections is negligible. Changes in the gain and time delay
quency content to a low-frequency mode associated with the second-
have a significant influence on the controller's effectiveness. As seen in
ary vortices.
Fig. 57, there are distinct regions in the λ τ parameter space for which
positive control effectiveness is obtained. With optimal combinations of

36
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 58. Closed loop control with λ ¼ 0:3, τ ¼ 30=36T0 and Cl0 ¼ 0:357 (Gao et al. [187]).

mechanical (passive) and fluidic (active) vortex generators in the sup-


pression of transonic shock oscillations. Through the production of co-
rotating vortices, vortex generators act to energise and stabilise the
boundary layer, promoting attached flow and inhibiting the shock-
induced separation that is instrumental to the buffet phenomenon
[190,191]. Caruana et al. [185,186] demonstrated the usefulness of
mechanical vortex generators experimentally on a rectangular OAT15A
wing section. A single configuration was considered by the authors,
Fig. 56. Control of three-dimensional buffet by closed-loop TED deflections (Caruana
consisting of rectangular vortex generators at a 30 incidence, with a
et al. [186]).
height equivalent to the local boundary layer height (h ¼ 1 mm) and
length l ¼ 5h. The vortex generators were positioned at 33% chord, 17%
chord upstream of the mean shock location, with a spanwise spacing of
λ ¼ 10h.
With the vortex generators installed an improvement in aerodynamic
performance is achieved at incidence close to buffet onset, however, drag
penalties are incurred at lower angles of attack. As shown in Fig. 59(a),
the additional energy imparted to the boundary layer by the vortex
generators acts to reduce the width of the wake, and correspondingly, the
drag coefficient at higher angles of attack. At conditions that exhibited
buffet in the clean wing configuration, the mechanical vortex generators
provide an effective means of suppression. From Fig. 59(b), the authors
show the low-frequency pressure fluctuations associated with large-scale
shock motions do not develop in the incidence range considered in the
experiments. Higher frequency fluctuations related to turbulent eddies in
separated flow regions are also attenuated at all test points.
Furthering the BUFET’N Co project, Dandois et al. [192] have con-
ducted a parametric study on the effects of various design parameters in
both mechanical and fluidic vortex generators. Numerical simulation of
transonic flow over a full-span constant cross-section OAT15A-CA aero-
foil at conditions with significant shock-induced separation was
employed to probe the parameter space. For the mechanical vortex
generators, the influence of skew angle and spanwise spacing (β and λ in
Fig. 57. Effective control regions for different λ and τ at M ¼ 0:70, α ¼ 5:5 and Re ¼ 3  Fig. 60, respectively) has been assessed, in addition to the chordwise
106 (Gao et al. [187]). positioning. The height and length of the vortex generators remains fixed
in the study, with h ¼ 1:3 mm and l ¼ 5h. Considering the lift and wall
7.2. Vortex generators pressure distributions as indicators of the parameters effectiveness, a
skew angle of β ¼ 0 is optimal, producing higher lift and a steady,
Under the guise of the BUFET’N Co project [182], a number of re- downstream shock location. In accordance with the previous work of
searchers at ONERA have further assessed the effectiveness of both Godard & Stanislas [193], the optimal spanwise spacing is found to be

37
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 61. Design parameters for fluidic vortex generators (Dandois et al. [192]).

1:3 mm, a spanwise spacing λ ¼ 10d and jet static pressure equal to the
local static pressure, yielding a momentum coefficient Cμ ¼ 2  103 .
The optimum skew angle exhibits a dependence on the pitch angle, with
higher pitch angles yielding a lower optimal skew angle. For the condi-
tions considered, the authors determine an optimal set of α ¼ 30 and β ¼
60 . With these optimal parameters, inserts were designed for a three-
dimensional half wing-body model to examine the behaviour of the
mechanical and fluidic vortex generators experimentally. Both methods
were effective in suppressing the large-scale separation present in the
clean wing configuration.
Molton et al. [194] and Dandois et al. [195] continued the work of
Dandois et al. [192] and detailed a comprehensive experimental inves-
tigation into the efficacy of both mechanical and fluidic vortex generators
on a swept half wing-body model. The reference configuration exhibits
large-scale separation at outboard sections, with established buffet at
M ¼ 0:82 and α ¼ 3:5 . In the clean configuration, the oil flow visual-
isations in Fig. 62 show the flow is fully separated downstream of the
shock past 50% span. Further, trailing edge pressure divergence, Kulite
pressure signals, PIV 2C and LDV 3C measurements all indicate sustained
shock oscillation at these conditions. For the controlled test cases, four
configurations have been investigated and are summarised in Table 3.
The test space consists of one mechanical configuration (VGm), two
Fig. 59. Effects of mechanical vortex generators on rectangular OAT15A wing section
(Caruana et al. [186]).
continuous fluidic vortex generators with varying skew (VGF4 and VGF5)
and one of pulsed fluidic actuation (VGFp). The oil traces in Fig. 62 show
the mechanical vortex generators suppress flow separation along the
majority of the span, excluding the region between 50 and 60% span

Fig. 60. Design parameters for mechanical vortex generators (Dandois et al. [192]).

λ ¼ 6h. Smaller spacing results in destructive interference between


adjacent vortices, reducing lift. Additionally, as long as the vortex gen-
erators are positioned upstream of the shock, the chordwise location has
a negligible influence on the aerodynamic performance.
Regarding the fluidic vortex generators, Dandois et al. [192] inves-
tigated the effects of varying pitch and skew angles (α and β in Fig. 61,
respectively). The orifice is positioned at 21% chord, with diameter d ¼ Fig. 62. Oil flow visualisations for the uncontrolled and controlled configurations: M ¼
0:82 and α ¼ 3:5 (Dandois et al. [195]).

38
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Table 3
Parameters for various vortex generator configurations [194].

Configuration Control type Dimensions (mm) λ βð Þ/LE Y=b (%) No. of VGs

1 Mechanical (VGm) h ¼ 1:3, l ¼ 5h 12h 0 51–89 27


2 Fluidic (VGF4) d¼1 6d 60 53–82 40
3 Fluidic (VGF5) d¼1 6d 30 53–82 40
4 Fluidic pulsed (VGFp) d¼1 11:5d 60 50–84 25

where a recirculation zone exists. A complex but near periodic three- mass flow rate for the continuous fluidic vortex generators shows that
dimensional shock/vortices interaction pattern is also present in the flow rate effects saturate for Cμ  5  104 . Regarding pulsed fluidic
shock foot region, with the vortices acting to deform the shock along the vortex generators, the forcing frequency has a profound influence on
span. The flow topology of the continuous fluidic vortex generators (not their effectiveness. As shown in Fig. 63(b), low-frequency pulsing results
shown) is strikingly similar, suppressing separation in the outboard re- in shock smearing, a characteristic of large-scale shock motion. At higher
gions and producing analogous shock/vortex interaction tracks. frequencies (Strouhal number St  0:24), the shock is stabilised and
Comparison of pressure distributions between the controlled and located marginally upstream relative to the continuous fluidic blowing.
uncontrolled cases (Fig. 63(a)) show that both mechanical and fluidic Molton et al. [194] demonstrated that mechanical, continuous and
control strategies shift the mean shock location downstream by approx- pulsed fluidic vortex generators in open loop are each effective in
imately 20% chord. The consistency of trailing edge pressures between delaying buffet onset by suppressing large-scale separation. Further,
the controlled cases, in addition to the RMS pressure distributions and fluidic blowing may be the preferable control strategy, providing similar
pressure spectra (not shown) confirm both control strategies effectively control actuation to mechanical vortex generators but with the absence of
damp out unsteadiness downstream of the shock. An analysis of varying a drag penalty at lower incidence.
Further evidence supporting the effectiveness of mechanical vortex
generators for buffet suppression has been provided by Timme & Sartor
[196]. Both steady and unsteady URANS simulations with the DLR-TAU
[61] code are performed on the half wing-body model RBC12, with ge-
ometry derived from their earlier investigations [153,154]. The simula-
tions are performed for various angles of attack at M ¼ 0:80 and Re ¼
3:75  106 . Co-rotating mechanical vortex generators are integrated at
32% chord between 64% and 91% span with spanwise spacing λ ¼ 7:7h,
taper ratio of 0.6 and aspect ratio of 1.3. From the steady simulations,
unconverged solutions are found at α ¼ 3 in the clean configuration and
α ¼ 3:6 for the controlled case. Unsteady simulations confirm that these
unconverged solutions are, in fact, due to the development of large-scale
unsteadiness and provide buffet onset conditions for the respective
configurations. The topological effects on the flow evidenced through
surface pressure distributions are in agreement with earlier studies
[111,194]. The vortex generators produce complex shock/vortex in-
teractions that shift the mean shock location downstream and inhibit the
development of massively separated flow. The integration of vortex
generators also significantly affects the frequency content of the aero-
dynamic coefficients. For the clean wing shown in Fig. 64(a), a distinct
peak is observed between 200 and 300 Hz near onset, which develops a
more broadband character with increasing incidence. For the controlled
wing in Fig. 64(b) however, low-amplitude, high-frequency content
dominates at onset. The buffet instability manifests at onset as a low-
amplitude broadband peak, which becomes dominant as the incidence
is increased.
Accompanying the recent surge in research of the three-dimensional
buffet mechanism, a number of publications from JAXA have further
explored the physics governing buffet suppression through mechanical
vortex generators. Kouchi et al. [197,198] have conducted wind tunnel
experiments to investigate shock motions on a two-dimensional NASA
SC(2)-0518 aerofoil with vortex generators. Focusing-Schlieren imaging
[199] and wavelet analysis revealed that although vortex generators
successfully suppressed onset to higher incidence, they significantly
increased the three-dimensionality of the flowfield and incited aperio-
dicity of the pressure fluctuations. Wavelet spectrograms of the time-
space flowfield maps further revealed that even in a suppressed buffet
state, frequency content relating to the low-frequency buffet mode and its
harmonics appears intermittently, with the rate of appearance dimin-
ishing with lower spanwise vortex generator spacing. Koike et al. [200]
have also investigated the differences between installation of vortex
generators on two- and three-dimensional wings experimentally on the
CRM, with the two-dimensional geometry derived from the 65% span
Fig. 63. Effects of mechanical and fluidic vortex generators on wall pressure distributions aerofoil section. Although not explicitly considering buffet conditions,
(Dandois et al. [195]).

39
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Raghunathan et al. [206] have provided a recent review on the appli-


cability of SCBs to the suppression of shock buffet. The findings reported
therein are not repeated here and the reader is directed to the work of
Raghunathan et al. [206] to gain context for the following discussion.
Eastwood & Jarrett [207] have performed a numerical investigation
using the commercial finite volume solver ANSYS Fluent [137], opti-
mising two- and three-dimensional SCBs for drag reduction and buffet
alleviation. The authors investigated the effects of spanwise width,
spacing and SCB edge design for three-dimensional bumps, in addition to
ramp angle, which is applicable in both the two- and three-dimensional
designs. The various design parameters are shown in Fig. 65. For design-
point aerodynamic improvements, two-dimensional SCBs outperform the
corresponding three-dimensional arrays. The usefulness of three-
dimensional SCBs arises in off-design conditions, where streamwise
vorticity is produced in a manner similar to vortex generators. By
reducing the length ltail or increasing the ramp angle θramp , a stronger
spanwise pressure gradient is achieved at the SCB tail, producing stron-
ger vortices which are more effective at suppressing the separated flow
downstream of the shock. Using a number of buffet onset metrics based
on wall shear stress from steady simulations, Eastwood & Jarrett [207]
ascertained that three-dimensional SCBs successfully reduced the large-
scale separation associated with buffet. Further, the buffet suppression
criteria exhibited a strong correlation to the design-point vortex strength,
permitting an assessment of both design-point aerodynamic performance
and improvements in buffet margin. These objectives are shown to be in
direct conflict, leaving the applicability of buffet alleviation through
SCBs subject to a trade-off with design-point performance.
A following study by Bogdanksi et al. [208] arrived at contrary con-
clusions on the usefulness of three-dimensional SCBs for buffet sup-
pression. Both steady and unsteady RANS simulations have been
performed with the DLR-TAU code [61] to investigate shock buffet
control on the aerofoil of the Pathfinder transonic laminar wing [209].
Three different SCB geometries are considered and shown in Fig. 66, with
a downhill simplex algorithm employed to optimise the height and
chordwise positioning for minimum drag at two design-points.
At both conditions, each of the SCBs degrades the buffet behaviour of
the aerofoil. As shown in Fig. 67, buffet onset occurs at an incidence
approximately 0:5 below the baseline in each of the controlled cases,
with higher buffet amplitudes observed within the buffet region. The
extended bump does provide an improvement in the separation behav-
iour of the aerofoil (Fig. 68), where the Hill-Shaped Control Bump

Fig. 64. Power spectral density of lift coefficient (Timme & Sartor [196]).

comparisons between the tests indicated the three-dimensional wing


exhibits greater sensitivity to installation effects. The subsequent paper
by Ito et al. [201], performing RANS computations on infinite-span three-
dimensional wings for various sweep angles, confirmed this sensitivity is
a consequence of cross-flow effects, which become more pronounced as
sweep increases.

7.3. Shock control bumps

In a similar manner to vortex generators, contoured shock control


bumps (SCBs) provide a passive means of shock control by modulating a
flowfield near the shock-wave/boundary layer interaction region to
reduce separation. The use of such devices for drag reduction is well
documented [202,203], with extensive studies into the technology con-
ducted under the EUROSHOCK and EUROSHOCK II projects [204,205]. Fig. 65. Three-dimensional SCB geometry (Eastwood & Jarrett [207]).

40
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

Fig. 66. SCB geometries investigated by Bogdanski et al. (Bogdanski et al. [208]).

Fig. 67. Lift polar for the three SCBs, M ¼ 0:76 (Bogdanski et al. [208]).

Fig. 68. Separated area over angle of attack for the three SCBs, M ¼ 0:76 (Bogdanski
(HSCB) and wedge designs perform poorly at off-design conditions.
et al. [208]).
Additionally, while the levels of separation increase with incidence, there
is no direct correlation to the lift variation, suggesting the amount of

41
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

separated flow present is a poor indicator of the performance of a buffet these various simulation parameters. The work of Goncalves & Houdeville
suppression technology. A similar conclusion was reached in evaluating [37] made significant progress in mapping the combinations of parameters
the influence of vortex strength on the buffet response. The three designs that successfully modelled the instability in a single test case. An extension
exhibit unique vortex structures, yet similar onset behaviour follows, of this work should be sought in future research efforts, comparing the
again indicative of a poor buffet onset metric. effective combinations of simulation parameters between various codes.
The subsequent study by Bogdanski et al. [210] further assessed the The growing applicability of hybrid RANS/LES methods to transonic
viability of various metrics as buffet indicators, in addition to evaluating buffet simulation has also been discussed. For capturing spectral data and
the performance of two innovative SCB geometries. A link is drawn be- the effects of secondary instabilities, such scale-resolving methods are the
tween the flow effects incurred through SCBs and classical vortex gen- only way forward from a computational perspective. Continued de-
erators, with SCBs generally able to produce half the streamwise vorticity velopments in the efficiency of such approaches is a necessity. Nonethe-
as relative to mechanical vortex generators. Correlations are also devel- less, the current requirements of high fidelity simulations to model the
oped between the levels of flow unsteadiness and the maximum lift co- buffet instability pose difficulties for the integration of buffet consider-
efficient, as well as a proposed buffet indicator: ations early in the aircraft design process. The modal nature of buffet
suggested by global stability theory may remedy this issue. Modal-based
α
BI ¼ ∫ αlow
high
⋅ClRMS dα (15) reduced order models, derived from methods such as proper orthogonal
decomposition or dynamic mode decomposition of the flowfield, could
where αlow is the highest incidence where all bumps exhibit steady flow provide a computationally efficient means of integrating buffet constraints
and αhigh is the lowest angle for which all bumps show lift oscillations. into aerofoil and wing optimisation.
Minimum coefficients of determination for the correlation of the A number of comprehensive experimental investigations have been
maximum lift and the buffet indicator to flow unsteadiness are 0.70 and detailed. The extensive database provided by the experimental buffet
0.66, respectively, indicating both these quantities serve as fair metrics campaign at ONERA, particularly for the OAT15A aerofoil, has been
for evaluating the efficacy of a buffet control technology. The authors indispensable for the validation of numerical simulations. These experi-
also investigated two forked bumps, and while the buffet response was ments have also provided a wealth of data for the development of a
marginally improved relative to previous designs, each of the SCBs promising theory of the mechanism governing buffet onset. To classify
exhibited degraded buffet behaviour relative to the baseline. definitively the mechanism of buffet onset, comprehensive experimental
studies must be performed on aerofoils of various geometry; encom-
8. Concluding remarks passing thin sections, classical NACA profiles and the inclusion of flap
and slat deflections. Such investigations will allow to research commu-
The phenomenon of self-sustained shock oscillation in the transonic nity to ascertain whether the mode of buffet exhibits any geometrical
regime is one that inhibits the performance of modern aircraft. Oscilla- dependency or whether the phenomenon is intrinsic to all aerofoils.
tory airloads degrade the handling characteristics of aircraft and the The understanding of dynamic interactions in the presence of shock
inherent structural vibrations associated with the unsteady flowfield buffet has progressed significantly in recent years. The exploration of the
have the potential to induce large amplitude limit cycle oscillations, aerodynamic resonance corresponding to buffet onset and the associated
degrading the structural integrity of a platform through cyclic loading. lock-in phenomenon determined by Raveh [125] offer a potential
Although over six decades has passed since the transonic buffet insta- mechanism for limit cycle oscillations in aircraft at transonic conditions.
bility was identified as a significant problem in aeronautics, a cohesive Additionally, research efforts conducted at Northwestern Polytechnical
explanation of the underlying mechanisms governing this phenomenon University indicate lock-in is not representative of a classical aero-
and robust buffet alleviation technologies continue to elude the dynamic resonance, but rather, an unstable coupling between structural
research community. and fluid modes. Although the phenomenon has been explicitly observed
This review has outlined the progress of recent research efforts in by Hartmann et al. [131] experimentally, the majority of investigations
furthering the understanding of the mechanism underlying buffet onset. in this field have been primarily concerned with two-dimensional ge-
Acoustic wave-propagation feedback models have resulted in excellent ometries. The work by Iovnovich et al. [134] provides a stepping stone
predictions of the buffet frequency in certain cases, however, evidence to for future research efforts in looking to confirm whether lock-in is a
the contrary has also been presented by a number of authors. The tran- realisable phenomenon in practical aircraft structures. Although costs
sonic pre-stall instability model has linked the onset of buffet to geo- will be high, experimental investigations will be imperative in this sense,
metric features of aerofoils, with an unstable shock-wave/separation supplemented by three-dimensional coupled Computational Fluid Dy-
bubble interaction driving the shock motion. A relationship between namics/Computational Structural Mechanics simulation with high-
buffet onset and a globally unstable aerodynamic mode has been estab- fidelity time-resolved or modal structural models.
lished, with support for this model from both experimental and numer- Research investigating the nature of three-dimensional shock buffet
ical studies steadily accumulating. To ascertain which, if any, of the has indicated that the predominantly two-dimensional phenomenon
currently prominent physical models of shock buffet correctly identifies observed on aerofoils and straight wings does not extend to swept-wing
physics governing the instability, future researchers should look to a geometries. For planforms consistent with modern civil transport
cohesive comparative study of predictions made by each of the models, aircraft, unsteady shock motions are first apparent at outboard sections
developed under a consistent framework and validated against a and progress inboard with increasing incidence. Frequency spectra of
comprehensive experimental database. pressure fluctuations on three-dimensional wings indicate the presence
The substantial body of research devoted to numerical simulation of of two broadband modes: the low-frequency shock oscillation that
the transonic shock oscillations illustrates that URANS methods are characterises buffet and a higher frequency bump representing a sec-
capable of capturing the low-frequency shock motions and intermittent ondary Kelvin-Helmholtz type instability. The reduced buffet frequencies
boundary layer separation inherent to shock buffet. In many instances, the pertaining to these three-dimensional geometries are significantly higher
buffet frequency, pressure fluctuations and velocity distributions predicted than those for equivalent two-dimensional aerofoils. While the three-
through simulation correlate well with experiment. A caveat of these dimensional buffet instability is yet to be completely understood,
positive results is that the simulation of shock buffet through URANS be- initial studies seem to indicate the presence of additional unstable
comes more so an art than a science. A demonstrable sensitivity to the aerodynamic modes, with the appearance of spanwise propagating
selected turbulence model, numerical scheme and spatial and temporal pressure disturbances linked to the lateral convection of buffet cells.
discretisation exists in the literature. The successful reproduction of the Exploration of the nature of autonomous shock oscillations on three-
transonic buffet phenomenon involves a complex interaction between dimensional geometries will be a significant topic in years to follow.

42
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

While the effects of sweep angle on the flow topology have been char- [17] Q. Xiao, H.M. Tsai, F. Liu, Numerical study of transonic buffet on a supercritical
airfoil, AIAA J. 44 (3) (2006) 620–628.
acterised numerically, the influence of taper, twist and other geometric
[18] S. Deck, Numerical simulation of transonic buffet over a supercritical airfoil, AIAA
parameters remain unknown. For future research efforts to gain a J. 43 (7) (2005) 1556–1566.
comprehensive understanding of geometric effects, any numerical [19] E. Garnier, S. Deck, Large-eddy simulation of transonic buffet over a supercritical
studies will need to be supplemented by experimental validation, for airfoil, in: Direct and Large-Eddy Simulation VII, ERCOFTAC Series, vol. 13,
Springer, Heidelberg, 2010, pp. 549–554.
which, the NASA Common Research Model offers an attractive test case. [20] L. Jacquin, P. Molton, S. Deck, B. Maury, D. Soulevant, Experimental study of
Finally, control of shock oscillations for two-dimensional sections has shock oscillation over a transonic supercritical profile, AIAA J. 47 (9) (2009)
been successfully achieved through closed-loop actuation of trailing edge 1985–1994.
[21] J.D. Crouch, A. Garbaruk, D. Magidov, Predicting the onset of flow unsteadiness
deflectors. Complexities have been observed when attempting to based on global instability, J. Comput. Phys. 224 (2) (2007) 924–940.
extrapolate this technology to three-dimensional wings. These diffi- [22] J.D. Crouch, A. Garbaruk, D. Magidov, L. Jacquin, Global structure of buffeting
culties may inherently be a result of the lack of understanding of the flow on transonic airfoils, in: IUTAM Symposium on Unsteady Separated Flows
and Their Control, Springer, 2009, pp. 297–306.
three-dimensional buffet phenomenon and exploration of the physics [23] P.L. Roe, Approximate Riemann solvers, parameter vectors, and difference
underlying this instability will aid the development of more effective schemes, J. Comput. Phys. 43 (2) (1981) 357–372.
control laws. Mechanical and fluidic vortex generators have also been [24] B. Van Leer, Upwind-difference methods for aerodynamic problems governed by
the Euler equations, Lect. Appl. Math. 22 (Part 2) (1985) 327–336.
shown to be effective buffet suppression technologies. Parametric studies [25] R.B. Lehoucq, D.C. Sorensen, C. Yang, ARPACK Users' Guide: Solution of Large-
of the design variables have led to a thorough understanding of vortex scale Eigenvalue Problems with Implicitly Restarted Arnoldi Methods, vol. 6,
generator configurations that are most effective at dispersing regions of Siam, 1998.
[26] J.B. McDevitt, A.F. Okuno, Static and Dynamic Pressure Measurements on a NACA
separated flow. Additionally, continual development in pulsed fluidic
0012 Airfoil in the Ames High Reynolds Number Facility, NASA TP-2485, National
vortex generators for buffet alleviation is expected. These configurations Aeronautics and Space Administration, 1985.
do not suffer the drag penalties incurred by mechanical vortex generators [27] A. Kuz’min, A. Shilkin, Transonic buffet over symmetric airfoils, in:
at cruise conditions and can be employed in closed-loop to yield more Computational Fluid Dynamics 2006, Springer, 2009, pp. 849–854.
[28] A. Kuz’min, Bifurcations and buffet of transonic flow past flattened surfaces,
effective control strategies. Relative to trailing edge deflections and Comput. Fluids 38 (7) (2009) 1369–1374.
vortex generators, the suitability of shock control bumps to mitigate [29] F. Sartor, C. Mettot, D. Sipp, Stability, receptivity, and sensitivity analyses of
shock oscillations has not been demonstrated. For this technology to buffeting transonic flow over a profile, AIAA J. 53 (7) (2014) 1980–1993.
[30] M. Iovnovich, D.E. Raveh, Reynolds-averaged Navier-Stokes study of the shock-
become fruitful, optimisation of the bump shapes and positions must buffet instability mechanism, AIAA J. 50 (4) (2012) 880–890.
explicitly consider a reliable indicator of buffet onset. [31] M. Adar, Y. Levy, A. Gany, Numerical simulation of flare safe separation, J. Aircr.
43 (4) (2006) 1129–1137.
[32] P.R. Spalart, S.R. Allmaras, A one-equation turbulence model for aerodynamic
Acknowledgements flows, La Rech. Aerosp. 1 (5) (1994).
[33] J.R. Edwards, S. Chandra, Comparison of eddy viscosity-transport turbulence
This research is partially funded by the Defence Science and Tech- models for three-dimensional, shock-separated flowfields, AIAA J. 34 (4) (1996)
756–763.
nology Group (MyIP: 6786), Australia. [34] L.L. Levy Jr., Experimental and computational steady and unsteady transonic
flows about a thick airfoil, AIAA J. 16 (6) (1978) 564–572.
References [35] J.W. Edwards, J.L. Thomas, Computational methods for unsteady transonic flows,
Unsteady Transonic Aerodyn. 120 (1989) 211–261.
[36] C.M. Maksymiuk, T.H. Pulliam, Viscous transonic airfoil workshop results using
[1] W.F. Hilton, R.G. Fowler, Photographs of Shock Wave Movement, NPL R&M No.
ARC2D, in: Proceedings of the 25th AIAA Aerospace Sciences Meeting, Reno, NV,
2692, National Physical Laboratories, 1947.
1987.
[2] D.G. Mabey, Oscillatory flows from shock induced separations on biconvex
[37] E. Goncalves, R. Houdeville, Turbulence model and numerical scheme assessment
aerofoils of varying thickness in ventilated wind tunnels, in: AGARD CP-296,
for buffet computations, Int. J. Numer. Methods Fluids 46 (11) (2004) 1127–1152.
Boundary layer effects on unsteady airloads, Aix-en-Provence, France, 14–19
[38] F. Grossi, M. Braza, Y. Hoarau, Prediction of transonic buffet by delayed detached-
September, 1981, p. 11, 1–14.
eddy simulation, AIAA J. 52 (10) (2014) 2300–2312.
[3] J. Gibb, The Cause and Cure of Periodic Flows at Transonic Speed (Ph.D. thesis),
[39] G. Barakos, D. Drikakis, Numerical simulation of transonic buffet flows using
Cranfield Institute of Tech., 1988.
various turbulence closures, Int. J. Heat Fluid Flow 21 (5) (2000) 620–626.
[4] D.G. Mabey, B.L. Welsh, B.E. Cripps, Periodic Flows on a Rigid 14% Thick
[40] G. Barakos, D. Drikakis, Implicit unfactored implementation of two-equation
Biconvex Wing at Transonic Speeds, RAE-TR-81059, British Royal Aircraft
turbulence models in compressible Navier-Stokes methods, Int. J. Numer. Methods
Establishment, 1981.
Fluids 28 (1) (1998) 73–94.
[5] H.H. Pearcey, Some Effects of Shock-induced Separation of Turbulent Boundary
[41] G. Barakos, D. Drikakis, An implicit unfactored method for unsteady turbulent
Layers in Transonic Flow Past Aerofoils, ARC R&M-3108, Aeronautical Research
compressible flows with moving boundaries, Comput. Fluids 28 (8) (1999)
Council, 1955.
899–922.
[6] H.H. Pearcey, A Method for the Prediction of the Onset of Buffeting and Other
[42] B.S. Baldwin, H. Lomax, Thin layer approximation and algebraic model for
Separation Effects from Wind Tunnel Tests on Rigid Models, AGARD TR 223,
separated turbulent flows, in: Proceedings of the 16th AIAA Aerospace Sciences
National Physics Laboratory, 1958.
Meeting, Hunstville, AL, 1978.
[7] H.H. Pearcey, D.W. Holder, Simple Methods for the Prediction of Wing Buffeting
[43] B.E. Launder, B.I. Sharma, Application of the energy-dissipation model of
Resulting from Bubble Type Separation, NPL AERO-REP-1024, National Physics
turbulence to the calculation of flow near a spinning disc, Lett. Heat Mass Transf. 1
Laboratory, 1962.
(2) (1974) 131–137.
[8] H.H. Pearcey, J. Osborne, A.B. Haines, The interaction between local effects at the
[44] Y. Nagano, C. Kim, A two-equation model for heat transport in wall turbulent
shock and rear separation-a source of significant scale effects in wind-tunnel tests
shear flows, J. Heat Transf. 110 (3) (1988) 583–589.
on aerofoils and wings, in: AGARD CP-35, Transonic Aerodynamics, Paris, France,
[45] D. Sofialidis, P. Prinos, Development of a non-linear, strain-sensitive k-ω
1968, p. 11, 1–23.
turbulence model, in: Proceedings of the 11th Symposium on Turbulent Shear
[9] J. Nitzsche, A numerical study on aerodynamic resonance in transonic separated
Flows, Grenoble, France, 1997, p. 2, 89–94.
flow, in: Proceedings of the International Forum on Aeroelasticity and Structural
[46] T.J. Craft, B.E. Launder, K. Suga, Development and application of a cubic eddy-
Dynamics, Seattle, WA, 2009.
viscosity model of turbulence, Int. J. Heat Fluid Flow 17 (2) (1996) 108–115.
[10] J.D. Crouch, A. Garbaruk, D. Magidov, A. Travin, Origin of transonic buffet on
[47] E. Goncalves, J.C. Robinet, R. Houdeville, Numerical simulation of transonic
aerofoils, J. Fluid Mech. 628 (2009) 357–369.
buffet over an airfoil, in: Proceedings of the 3rd International Symposium on
[11] H. Tijdeman, Investigations of the Transonic Flow Around Oscillating Airfoils
Turbulence and Shear Flow Phenomena, Senda, Japan, 2003.
(PhD Thesis), Delft University of Technology, TU Delft, 1977.
[48] V. Couaillier, Numerical simulation of separated turbulent flows based on the
[12] B.H.K. Lee, Self-sustained shock oscillations on airfoils at transonic speeds, Prog.
solution of RANS/low Reynolds two-equation model, in: Proceedings of the 37th
Aerosp. Sci. 37 (2) (2001) 147–196.
Aerospace Sciences Meeting and Exhibit, Reno, NV, 1999.
[13] B.H.K. Lee, Oscillatory shock motion caused by transonic shock boundary-layer
[49] B. Smith, The k-kl turbulence model and wall layer model for compressible flows,
interaction, AIAA J. 28 (5) (1990) 942–944.
in: Proceedings of the 21st Fluid Dynamics, Plasma Dynamics and Lasers
[14] S. Raghunathan, R.D. Mitchell, M.A. Gillan, Transonic shock oscillations on
Conference, Seattle, WA, 1990.
NACA0012 aerofoil, Shock Waves 8 (4) (1998) 191–202.
[50] B. Smith, A near wall model for the k-l two equation turbulence model, in:
[15] B.H.K. Lee, Investigation of flow separation on a supercritical airfoil, J. Aircr. 26
Proceedings of the 24th AIAA Fluid Dynamics Conference, Colorado Springs, CO,
(11) (1989) 1032–1037.
1994.
[16] A.L. Erickson, J.D. Stephenson, A Suggested Method of Analyzing Transonic
[51] D.C. Wilcox, Reassessment of the scale-determining equation for advanced
Flutter of Control Surfaces Based on Available Experimental Evidence, NACA RM
turbulence models, AIAA J. 26 (11) (1988) 1299–1310.
A7F30, National Advisory Committee for Aeronautics, 1947.

43
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

[52] F.R. Menter, Two-equation eddy-viscosity turbulence models for engineering [85] G. Barbut, M. Braza, Y. Hoarau, G. Barakos, A. Sevrain, J.B. Vos, Prediction of
applications, AIAA J. 32 (8) (1994) 1598–1605. transonic buffet around a wing with flap, in: Progress in Hybrid RANS-LES
[53] J.C. Kok, Resolving the dependence on freestream values for the k-turbulence Modelling, Springer, 2010, pp. 191–204.
model, AIAA J. 38 (7) (2000) 1292–1295. [86] M. Braza, NACA0012 with aileron, in: P. Doerffer, C. Hirsch, J.P. Dussauge,
[54] W.P. Jones, B.E. Launder, The prediction of laminarization with a two-equation H. Babinsky, G.N. Barakos (Eds.), Unsteady Effects of Shock Wave Induced
model of turbulence, Int. J. Heat Mass Transf. 15 (2) (1972) 301–314. Separation, Springer, 2010, pp. 101–131.
[55] A. Harten, B. Engquist, S. Osher, S.R. Chakravarthy, Uniformly high order accurate [87] I. Mary, P. Sagaut, Large eddy simulation of flow around an airfoil near stall, AIAA
essentially non-oscillatory schemes, III, J. Comput. Phys. 71 (2) (1987) 231–303. J. 40 (6) (2002) 1139–1145.
[56] M. Thiery, E. Coustols, URANS computations of shock-induced oscillations over [88] M. Pechier, Previsions numeriques de l’effet magnus pour des configurations de
2D rigid airfoils: influence of test section geometry, Flow, Turbul. Combust. 74 (4) munitions (Ph.D. thesis), 1999.
(2005) 331–354. [89] P. Spalart, W. Jou, M. Strelets, S. Allmaras, et al., Comments on the feasibility of
[57] L. Cambier, M. Gazaix, elsA: an efficient object-oriented solution to CFD les for wings, and on a hybrid rans/les approach, Adv. DNS/LES 1 (1997) 4–8.
complexity, in: Proceedings of the 40th AIAA Aerospace Sciences Meeting & [90] B. Raverdy, I. Mary, P. Sagaut, N. Liamis, High-resolution large-eddy simulation of
Exhibit, Reno, NV, 2002. flow around low-pressure turbine blade, AIAA J. 41 (3) (2003) 390–397.
[58] A. Jameson, W. Schmidt, E. Turkel, Numerical solution of the Euler equations by [91] S.M. Kay, S.L. Marple, Spectrum analysis - a modern perspective, Proc. IEEE 69
finite volume methods using Runge Kutta time stepping schemes, in: Proceedings (11) (1981) 1380–1419.
of the 14th AIAA Fluid and Plasma Dynamics Conference, Palo Alto, CA, 1981. [92] L. Larchev^eque, P. Sagaut, T.H. Le, P. Comte, Large-eddy simulation of a
[59] J. Cousteix, V. Saint-Martin, R. Messing, H. Bezard, B. Aupoix, Development of the compressible flow in a three-dimensional open cavity at high Reynolds number,
k-ϕ turbulence model, in: 11th Symposium on Turbulent Shear Flows, Institut J. Fluid Mech. 516 (2004) 265–301.
National Polytechnique, Universite Joseph Fourier, Grenoble, France, 1997. [93] F. Ducros, V. Ferrand, F. Nicoud, C. Weber, D. Darracq, C. Gacherieu, et al., Large-
[60] S. Illi, T. Lutz, E. Kr€
amer, On the capability of unsteady RANS to predict transonic eddy simulation of the shock/turbulence interaction, J. Comput. Phys. 152 (2)
buffet, in: Proceedings of the Third Symposium Simulation of Wing and Nacelle (1999) 517–549.
Stall; Braunschweig, Germany, 2012. [94] E. Lenormand, P. Sagaut, L.T. Phuoc, P. Comte, Subgrid-scale models for large-
[61] D. Schwamborn, T. Gerhold, R. Heinrich, The DLR TAU-Code: recent applications eddy simulations of compressible wall bounded flows, AIAA J. 38 (8) (2000)
in research and industry, in: Proceedings of the European Conference on 1340–1350.
Computational Fluid Dynamics ECCOMAS CFD 2006; Egmond Aan Zee, The [95] J.B. Vos, A. Rizzi, A. Corjon, E. Chaput, E. Soinne, Recent advances in
Netherlands, 2006. aerodynamics inside the NSMB (Navier-Stokes Multi-Block) consortium, in:
[62] T. Rung, U. Bunge, M. Schatz, F. Thiele, Restatement of the Spalart-Allmaras eddy- Proceedings of the 36th AIAA Aerospace Sciences Meeting & Exhibit, Reno, NV,
viscosity model in strain-adaptive formulation, AIAA J. 41 (7) (2003) 1396–1399. 1998.
[63] T. Rung, H. Lübcke, M. Franke, L. Xue, F. Thiele, S. Fu, Assessment of explicit [96] F. Grossi, M. Braza, Y. Hoarau, Delayed Detached-Eddy Simulation of the
algebraic stress models in transonic flows, Eng. Turbul. Model. Exp. 4 (1999) transonic flow around a supercritical airfoil in the buffet regime, in: Progress in
659–668. Hybrid RANS-LES Modelling, Springer, 2012, pp. 369–378.
[64] S. Jakirlic, K. Hanjalic, A new approach to modelling near-wall turbulence energy [97] F. Grossi, Physics and Modeling of Unsteady Shock Wave/boundary Layer
and stress dissipation, J. Fluid Mech. 459 (2002) 139–166. Interactions over Transonic Airfoils by Numerical Simulation (Ph.D. thesis),
[65] A. Kourta, G. Petit, J.C. Courty, J.P. Rosenblum, Buffeting in transonic flow National Polytechnic Institute of Toulouse, 2014.
prediction using time-dependent turbulence model, Int. J. Numer. Methods Fluids [98] B. Roidl, M. Meinke, W. Schr€ oder, Zonal RANS-LES computation of transonic
49 (2) (2005) 171–182. airfoil flow, in: Proceedings of the 29th AIAA Applied Aerodynamics Conference,
[66] A. Kourta, Computation of vortex shedding in solid rocket motors using time- Honolulu, Hawaii, 2011.
dependent turbulence model, J. Propuls. Power 15 (3) (1999) 390–400. [99] J.L. Fulker, M.J. Simmons, An experimental investigation of passive shock/
[67] T.H. Shih, J. Zhu, J.L. Lumley, A new Reynolds stress algebraic equation model, boundary-layer control on an aerofoil, in: EUROSHOCK-Drag Reduction by
Comput. Methods Appl. Mech. Eng. 125 (1–4) (1995) 287–302. Passive Shock Control, Springer, 1997, pp. 379–400.
[68] V. Brunet, Computational study of buffet phenomenon with unsteady RANS [100] M.S. Liou, C.J. Steffen, A new flux splitting scheme, J. Comput. Phys. 107 (1)
equations, in: Proceedings of the 21st AIAA Applied Aerodynamics Conference, (1993) 23–39.
Orlando, FL, 2003. [101] J.P. Boris, F.F. Grinstein, E.S. Oran, R.L. Kolbe, New insights into large-eddy
[69] A. Soda, N. Verdon, Investigation of influence of different modelling parameters simulation, Fluid Dyn. Res. 10 (4–6) (1992) 199–228.
on calculation of transonic buffet phenomena, in: New Results in Numerical and [102] Q. Zhang, W. Schr€ oder, M. Meinke, A zonal RANS-LES method to determine the
Experimental Fluid Mechanics V, Springer, 2006, pp. 487–495. flow over a high-lift configuration, Comput. Fluids 39 (7) (2010) 1241–1253.
[70] M.E. Olsen, T.J. Coakley, The lag model, a turbulence model for non equilibrium [103] J.B. McDevitt, L.L. Levy Jr., G.S. Deiwert, Transonic flow about a thick circular-arc
flows, in: Proceedings of the 15th AIAA Computational Fluid Dynamics airfoil, AIAA J. 14 (5) (1976) 606–613.
Conference, Anaheim, CA, 2001. [104] J.B. McDevitt, Supercritical Flow about a Thick Circular-arc Airfoil, NASA-TM-
[71] A.B.M.T. Hasan, M.M. Alam, RANS computation of transonic buffet over a 78549, National Aeronautics and Space Administration, 1979.
supercritical airfoil, Procedia Eng. 56 (2013) 303–309. [105] K. Finke, Unsteady shock wave-boundary layer interaction on profiles in transonic
[72] A.A. Rokoni, A.B.M.T. Hasan, Prediction of transonic buffet onset for flow over a flow, in: AGARD CP-183, Flow Separation, Gottingen, Germany, 27–30 May,
supercritical airfoil-a numerical investigation, J. Mech. Eng. 43 (1) (2013) 48–53. 1975, p. 28, 1–11.
[73] R. Carrese, P. Marzocca, O. Levinski, N. Joseph, Investigation of supercritical [106] E. Stanewsky, D. Basler, Experimental investigation of buffet onset and
airfoil dynamic response due to transonic buffet, in: Proceedings of the 54th AIAA penetration on a supercritical airfoil at transonic speeds, in: AGARD CP-483,
Aerospace Sciences Meeting, San Diego, CA, 2016. Aircraft Dynamic Loads Due to Flow Separation, Sorrento, Italy, 1990, p. 4, 1–11.
[74] N.F. Giannelis, G.A. Vio, Aeroelastic interactions of a supercritical aerofoil in the [107] F.W. Roos, Surface pressure and wake flow fluctuations in a supercritical airfoil
presence of transonic shock buffet, in: Proceedings of the 54th AIAA Aerospace flow field, in: Proceedings of the 13th AIAA Aerospace Sciences Meeting,
Sciences Meeting, San Diego, CA, 2016. Pasadena, CA, 1975.
[75] B.E. Launder, G. Reece Jr., W. Rodi, Progress in the development of a Reynolds- [108] F.W. Roos, Some features of the unsteady pressure field in transonic airfoil
stress turbulence closure, J. Fluid Mech. 68 (03) (1975) 537–566. buffeting, J. Aircr. 17 (11) (1980) 781–788.
[76] B. Van Leer, Towards the ultimate conservative difference scheme. V. A second- [109] N. Hirose, H. Miwa, Computational and experimental research on buffet
order sequel to Godunov's method, J. Comput. Phys. 32 (1) (1979) 101–136. phenomena of transonic airfoils, in: Symposium Transsonicum III, Springer, 1989,
[77] M.S. Liou, A sequel to AUSM: AUSMþ, J. Comput. Phys. 129 (2) (1996) 364–382. pp. 489–499.
[78] J.T. Xiong, Y. Liu, F. Liu, S. Luo, Z. Zhao, X. Ren, et al., Multiple solutions and [110] B.H.K. Lee, F.A. Ellis, J. Bureau, Investigation of the buffet characteristics of two
buffet of transonic flow over NACA0012 airfoil, in: 31st AIAA Applied supercritical airfoils, J. Aircr. 26 (8) (1989) 731–736.
Aerodynamics Conference, 2013. [111] J. Dandois, P. Molton, A. Lepage, A. Geeraert, V. Brunet, J.B. Dor, et al., Buffet
[79] V. Brunet, S. Deck, P. Molton, M. Thiery, A complete experimental and numerical characterization and control for turbulent wings, AerospaceLab 6 (2013) 1–17.
study of the buffet phenomenon over the OAT15A airfoil, ONERA Tire Part 35 [112] L. Jacquin, P. Molton, S. Deck, B. Maury, D. Soulevant, An experimental study of
(2005) 1–9. shock oscillation over a transonic supercritical profile, in: Proceedings of the 35th
[80] O. Rouzaud, S. Plot, V. Couaillier, Numerical simulation of buffeting over airfoil AIAA Fluid Dynamics Conference and Exhibit, Toronto, Ont, 2005.
using dual time stepping method, in: Proceedings of the European Conference on [113] A. Hartmann, M. Klaas, W. Schr€ oder, Time-resolved stereo PIV measurements of
Computational Fluid Dynamics ECCOMAS CFD 2000; Barcelona, Spain, 2000. shock-boundary layer interaction on a supercritical airfoil, Exp. Fluids 52 (3)
[81] C.L. Rumsey, M.D. Sanetrik, R.T. Biedron, N.D. Melson, E.B. Parlette, Efficiency (2012) 591–604.
and accuracy of time-accurate turbulent Navier-Stokes computations, Comput. [114] A. Hartmann, A. Feldhusen, W. Schr€ oder, On the interaction of shock waves and
Fluids 25 (2) (1996) 217–236. sound waves in transonic buffet flow, Phys. Fluids (1994–present) 25 (2) (2013)
[82] F. Furlano, E. Coustols, S. Plot, O. Rouzaud, Steady and unsteady computations of 026101.
flows close to airfoil buffeting: validation of turbulence models, in: Proceedings of [115] B.H.K. Lee, Transonic buffet on a supercritical aerofoil, Aeronaut. J. 94 (935)
the Turbulence and Shear Flow Phenomena Conference, Stockholm, Sweden, (1990) 143–152.
2001, pp. 211–216. [116] M.S. Howe, Trailing edge noise at low Mach numbers, J. Sound Vib. 225 (2)
[83] A.M. Vuillot, V. Couaillier, N. Liamis, 3D turbomachinery Euler and Navier-Stokes (1999) 211–238.
calculations with a multidomain cell centered approach, in: Proceedings of the [117] R. Ewert, W. Schr€ oder, On the simulation of trailing edge noise with a hybrid LES/
29th Joint Propulsion Conference and Exhibit, Monterey, CA, 1993. APE method, J. Sound Vib. 270 (3) (2004) 509–524.
[84] A. Garbaruk, M. Shur, M. Strelets, P.R. Spalart, Numerical study of wind-tunnel
walls effects on transonic airfoil flow, AIAA J. 41 (6) (2003) 1046–1054.

44
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

[118] A. Alshabu, H. Olivier, I. Klioutchnikov, Investigation of upstream moving [150] B. Benoit, I. Legrain, Buffeting prediction for transport aircraft applications based
pressure waves on a supercritical airfoil, Aerosp. Sci. Technol. 10 (6) (2006) on unsteady pressure measurements, in: Proceedings of the 5th AIAA Applied
465–473. Aerodynamics Conference, Monterey, CA, 1987, pp. 225–235.
[119] A. Alshabu, H. Olivier, Unsteady wave phenomena on a supercritical airfoil, AIAA [151] P.C. Steimle, D.C. Karhoff, W. Schr€ oder, Unsteady transonic flow over a transport-
J. 46 (8) (2008) 2066–2073. type swept wing, AIAA J. 50 (2) (2012) 399–415.
[120] R.J. Zwaan, NLR 7301 Supercritical Airfoil Oscillatory Pitching and Oscillating [152] V. Brunet, S. Deck, Zonal-detached eddy simulation of transonic buffet on a civil
Flap. Compendium of Unsteady Aerodynamics Measurements AGARD-R-702, aircraft type configuration, in: Advances in Hybrid RANS-LES Modelling, Springer,
Advisory Group for Aerospace Research and Development, 1982. 2008, pp. 182–191.
[121] S.S. Davis, NACA 64A010 Oscillatory Pitching. Compendium of Unsteady [153] F. Sartor, S. Timme, Reynolds-Averaged Navier-Stokes simulations of shock buffet
Aerodynamics Measurements AGARD-R-702, Advisory Group for Aerospace on half wing-body configuration, in: Proceedings of the 53rd AIAA Aerospace
Research and Development, 1982. Sciences Meeting, Kissimmee, FL, 2015.
[122] R.H. Landon, NACA 0012 Oscillatory and Transient Pitching. Compendium of [154] F. Sartor, S. Timme, Delayed detached-eddy simulation of shock buffet on half
Unsteady Aerodynamics Measurements AGARD-R-702, Advisory Group for wing-body configuration, in: Proceedings of the 22nd AIAA Computational Fluid
Aerospace Research and Development, 1982. Dynamics Conference, Dallas, TX, 2015.
[123] S.S. Davis, G.N. Malcolm, Transonic shock-wave/boundary-layer interactions on [155] F. Sartor, S. Timme, Delayed detached-eddy simulation of shock buffet on half
an oscillating airfoil, AIAA J. 18 (11) (1980) 1306–1312. wing-body configuration, AIAA J. 55 (4) (2017) 1230–1240.
[124] C. Despere, D. Caruanal, A. Mignosi, O. Reberga, M. Correge, H. Gassot, et al., [156] S.G. Lawson, D. Greenwell, M. Quinn, Characterisation of buffet on a civil aircraft
Buffet active control-experimental and numerical results, in: Proceedings of RTO wing, in: Proceedings of the 54th AIAA Aerospace Sciences Meeting, San Diego,
AVT Symposium on Active Control Technology of Enhanced Performance CA, 2016.
Operational Capabilities of Military Aircraft, Land Vehicles, and Sea Vehicles; [157] J. Peiro, J. Peraire, K. Morgan, O. Hassan, N. Birch, The numerical simulation of
Braunschweig, Germany, 2001. flow about installed aero engine nacelle using a finite element Euler solver on
[125] D.E. Raveh, Numerical study of an oscillating airfoil in transonic buffeting flows, unstructured meshes, Aeronaut. J. 96 (956) (1992) 224–232.
AIAA J. 47 (3) (2009) 505–515. [158] ESDU, An Introduction to Aircraft Buffet and Buffeting, ESDU 87012, ESDU, July
[126] D.C. Wilcox, Formulation of the kw turbulence model revisited, AIAA J. 46 (11) 1987.
(2008) 2823–2838. [159] L. Masini, S. Timme, A. Ciarella, A. Peace, Influence of vane vortex generators on
[127] E.H. Dowell, K.C. Hall, J.P. Thomas, R.E. Kielb, M.A. Spiker, C.M. Denegri Jr., transonic wing buffet: further analysis of the BUCOLIC experimental dataset, in:
A new solution method for unsteady flows around oscillating bluff bodies, in: Proceedings of the 52nd 3AF International Conference on Applied Aerodynamics,
IUTAM Symposium on Fluid-structure Interaction in Ocean Engineering, Springer, Lyon, France, 2017.
2008, pp. 37–44. [160] J. Dandois, Experimental study of transonic buffet phenomenon on a 3D swept
[128] D.E. Raveh, E.H. Dowell, Frequency lock-in phenomenon for oscillating airfoils in wing, Phys. Fluids 28 (1) (2016) 016101.
buffeting flows, J. Fluids Struct. 27 (1) (2011) 89–104. [161] M. Iovnovich, D.E. Raveh, Numerical study of shock buffet on three-dimensional
[129] N.F. Giannelis, G.A. Vio, Investigation of frequency lock-in phenomena on a wings, AIAA J. 53 (2) (2014) 449–463.
supercritical aerofoil in the presence of transonic shock oscillations, in: [162] J. Vassberg, M. Dehaan, M. Rivers, R. Wahls, Development of a common research
Proceedings of the 17th International Forum on Aeroelasticity and Structural model for applied CFD validation studies, in: Proceedings of the 26th AIAA
Dynamics, Como, Italy, 2017. Applied Aerodynamics Conference, Honolulu, Hawaii, 2008.
[130] M. Iovnovich, D.E. Raveh, Transonic unsteady aerodynamics in the vicinity of [163] S. Koike, M. Ueno, K. Nakakita, A. Hashimoto, Unsteady pressure measurement of
shock-buffet instability, J. Fluids Struct. 29 (2012) 131–142. transonic buffet on NASA Common Research Model, in: Proceedings of the 34th
[131] A. Hartmann, M. Klaas, W. Schr€ oder, Coupled airfoil heave/pitch oscillations at AIAA Applied Aerodynamics Conference, Washington, D.C, 2016.
buffet flow, AIAA J. 51 (7) (2013) 1542–1552. [164] M. Ueno, M. Kohzai, S. Koga, H. Kato, K. Nakakita, N. Sudani, 80% scaled NASA
[132] G. Dietz, G. Schewe, H. Mai, Amplification and amplitude limitation of heave/ Common Research Model wind tunnel test of JAXA at relatively low Reynolds
pitch limit-cycle oscillations close to the transonic dip, J. Fluids Struct. 22 (4) number, in: Proceedings of the 51st AIAA Aerospace Sciences Meeting, Grapevine,
(2006) 505–527. TX, 2013.
[133] D.E. Raveh, E.H. Dowell, Aeroelastic responses of elastically suspended airfoil [165] Y. Sugioka, D. Numata, K. Asai, S. Koike, K. Nakakita, S. Koga, Unsteady PSP
systems in transonic buffeting flows, AIAA J. 52 (5) (2014) 926–934. measurement of transonic buffet on a wing, in: Proceedings of the 53rd AIAA
[134] M. Iovnovich, D. Michaels, M. Adar, D.E. Raveh, Numerical study of the transonic Aerospace Sciences Meeting, Kissimmee, FL, 2015.
limit cycle oscillation phenomenon on the F-16 fighter aircraft, in: Proceedings of [166] D. Caruana, A. Mignosi, M. Correge, A. Le Pourhiet, Buffeting active control in
the 16th International Forum on Aeroelasticity and Structural Dynamics, St transonic flow, in: Proceedings of the 21st AIAA Applied Aerodynamics
Petersburg, Russia, 2015. Conference, Orlando, FL, 2003.
[135] J. Quan, W. Zhang, C. Gao, Z. Ye, Characteristic analysis of lock-in for an [167] F. Sartor, S. Timme, Mach number effects on buffeting flow on a half wing-body
elastically suspended airfoil in transonic buffet flow, Chin. J. Aeronaut. 29 (1) configuration, Int. J. Numer. Methods Heat Fluid Flow 26 (7) (2016) 2066–2080.
(2016) 129–143. [168] Y. Ohmichi, A. Hashimoto, Numerical investigation of transonic buffet on a three-
[136] N.F. Giannelis, G.A. Vio, G. Dimitriadis, Dynamic interactions of a supercritical dimensional wing using incremental mode decomposition, in: Proceedings of the
aerofoil in the presence of transonic shock buffet, in: Proceedings of the 27th 55th AIAA Aerospace Sciences Meeting, Grapevine, TX, 2017.
International Conference on Noise and Vibration Engineering, Leuven, Belgium, [169] S. Timme, R. Thormann, Towards three-dimensional global stability analysis of
2016. transonic shock buffet, in: Proceedings of the AIAA Atmospheric Flight Mechanics
[137] ANSYS, Fluent 12. 1 User Manual, ANSYS Inc, 2011. Conference, Washington, D.C, 2016.
[138] N.F. Giannelis, J.A. Geoghegan, G.A. Vio, Gust response of a supercritical aerofoil [170] Y. Saad, M.H. Schultz, GMRES: a generalized minimal residual algorithm for
in the vicinity of transonic shock buffet, in: Proceedings of the 20th Australasian solving nonsymmetric linear systems, SIAM J. Sci. Stat. Comput. 7 (3) (1986)
Fluid Mechanics Conference, Perth, Western Australia, 2016. 856–869.
[139] E.H. Dowell, J. Edwards, T. Strganac, Nonlinear aeroelasticity, J. Aircr. 40 (5) [171] S. Xu, S. Timme, K.J. Badcock, Krylov subspace recycling for linearised
(2003) 857–874. aerodynamics analysis using DLR-TAU, in: Proceedings of the 16th International
[140] M.J.d.C. Henshaw, K.J. Badcock, G.A. Vio, C.B. Allen, J. Chamberlain, I. Kaynes, et Forum on Aeroelasticity and Structural Dynamics, St Petersburg, Russia, 2015.
al., Non-linear aeroelastic prediction for aircraft applications, Prog. Aerosp. Sci. 43 [172] J. Xiong, F. Liu, Numerical simulation of transonic buffet on swept wing of
(4) (2007) 65–137. supercritical airfoils, in: Proceedings of the 543rd AIAA Fluid Dynamics
[141] O.O. Bendiksen, Review of unsteady transonic aerodynamics: theory and Conference, San Diego, CA, 2013.
applications, Prog. Aerosp. Sci. 47 (2) (2011) 135–167. [173] S.A. Illi, C. Fingskes, T. Lutz, E. Kr€amer, Transonic tail buffet simulations for the
[142] C. Gao, W. Zhang, Y. Liu, Z. Ye, Y. Jiang, Numerical study on the correlation of common research model, in: Proceedings of the 31st AIAA Applied Aerodynamics
transonic single-degree-of-freedom flutter and buffet, Sci. China Phys. Mech. Conference, San Diego, CA, 2013.
Astron. 58 (8) (2015) 1–12. [174] S.A. Illi, T. Lutz, E. Kr€amer, Transonic tail buffet simulations on the ATRA research
[143] C. Gao, W. Zhang, Z. Ye, A new viewpoint on the mechanism of transonic single- aircraft, in: Computational Flight Testing, Springer, 2013, pp. 273–287.
degree-of-freedom flutter, Aerosp. Sci. Technol. 52 (2016) 144–156. [175] A.F. Ribeiro, B. Konig, D. Singh, E. Fares, R. Zhang, P. Gopalakrishnan, et al.,
[144] W. Zhang, C. Gao, Y. Liu, Z. Ye, Y. Jiang, The interaction between flutter and Buffet simulations with a Lattice-Boltzmann based transonic solver, in:
buffet in transonic flow, Nonlinear Dyn. 82 (4) (2015) 1851–1865. Proceedings of the 55th AIAA Aerospace Sciences Meeting, Grapevine, TX, 2017.
[145] J.A. Rivera, B.E. Dansberry, R.M. Bennett, M.H. Durham, W.A. Silva, NACA0012 [176] D.P. Lockard, L.S. Luo, S.D. Milder, B.A. Singer, Evaluation of PowerFLOW for
Benchmark Model Experimental Flutter Results with Unsteady Pressure aerodynamic applications, J. Stat. Phys. 107 (1–2) (2002) 423–478.
Distributions, National Aeronautics and Space Administration, 1992. [177] A. Hashimoto, K. Murakami, T. Aoyama, K. Ishiko, M. Hishida, M. Sakashita, et al.,
[146] C. Gao, W. Zhang, X. Li, Y. Liu, J. Quan, Z. Ye, et al., Mechanism of frequency lock- Toward the fastest unstructured CFD code ‘FaSTAR’, in: Proceedings of the 50th
in in transonic buffeting flow, J. Fluid Mech. 818 (2017) 528–561. AIAA Aerospace Sciences Meeting Including the New Horizons Forum and
[147] N.C. Perkins, C.D. Mote, Comments on curve veering in eigenvalue problems, Aerospace Exposition, Nashville, TN, 2012.
J. Sound Vib. 106 (3) (1986) 451–463. [178] T. Ishida, A. Hashimoto, Y. Ohmichi, T. Aoyama, K. Takekawa, Transonic buffet
[148] P. Meliga, J.M. Chomaz, An asymptotic expansion for the vortex-induced simulation over NASA-CRM by Unsteady-FaSTAR code, in: Proceedings of the 55th
vibrations of a circular cylinder, J. Fluid Mech. 671 (2011) 137–167. AIAA Aerospace Sciences Meeting, Grapevine, TX, 2017.
[149] F. Roos, The buffeting pressure field of a high-aspect-ratio swept wing, in: [179] G.K. Kenway, J. Martins, Aerodynamic shape optimization of the CRM
Proceedings of the 18th Fluid Dynamics and Plasmadynamics and Lasers configuration including buffet-onset conditions, in: Proceedings of the 54th AIAA
Conference, Cincinatti, OH, 1985. Aerospace Sciences Meeting, San Diego, CA, 2016.

45
N.F. Giannelis et al. Progress in Aerospace Sciences xxx (2017) 1–46

[180] B.H.K. Lee, Effects of trailing-edge flap on buffet characteristics of a supercritical [197] T. Kouchi, S. Yamaguchi, S. Koike, T. Nakajima, M. Sato, H. Kanda, et al., Wavelet
airfoil, J. Aircr. 29 (1) (1992) 93–100. analysis of transonic buffet on a two-dimensional airfoil with vortex generators,
[181] B.H.K. Lee, F.C. Tang, Transonic buffet of a supercritical airfoil with trailing-edge Exp. Fluids 57 (11) (2016) 166.
flap, J. Aircr. 26 (5) (1989) 459–464. [198] T. Kouchi, S. Yamaguchi, S. Koike, S. Yanase, T. Nakajima, M. Sato, et al., Wavelet
[182] E. Coustols, V. Brunet, R. Bur, D. Caruana, D. Sipp, BUFETN Co: a joint ONERA analysis of unsteady shock-wave motion on two-dimensional airfoil with vortex
research project devoted to buffet control on a transonic 3D wing using a closed- generators, in: Proceedings of the 54th AIAA Aerospace Sciences Meeting, San
loop approach, in: Proceedings of the CEAS/KATNET II Conference; Bremen, Diego, CA, 2016.
Germany, 2009. [199] T. Kouchi, C. Goyne, R. Rockwell, R. Reynolds, R. Krauss, J. McDaniel, Focusing-
[183] J.C. Le Balleur, P. Girodroux-Lavigne, A semi-implicit and unsteady numerical Schlieren visualization in direct-connect dual-mode scramjet, in: Proceedings of
method of viscous-inviscid interaction for transonic separated flows, La Rech. the 18th AIAA/3AF International Space Planes and Hypersonic Systems and
Aerosp. 1984 (1) (1984) 15–37. Technologies Conference, Tours, France, 2012.
[184] J.C. Le Balleur, P. Girodroux-Lavigne, A viscousinviscid interaction method for [200] S. Koike, K. Nakakita, T. Nakajima, S. Koga, M. Sato, H. Kanda, et al., Experimental
computing unsteady transonic separation, in: Numerical and Physical Aspects of investigation of vortex generator effect on two- and three-dimensional NASA
Aerodynamic Flows III, Springer, 1986, pp. 252–271. common research models, in: Proceedings of the 53rd AIAA Aerospace Sciences
[185] D. Caruana, A. Mignosi, C. Robitaillie, M. Correge, Separated flow and buffeting Meeting, Kissimmee, FL, 2015.
control, Flow, Turbul. Combust. 71 (1–4) (2003) 221–245. [201] Y. Ito, K. Yamamoto, K. Kusunose, S. Koike, K. Nakakita, M. Murayama, et al.,
[186] D. Caruana, A. Mignosi, M. Correge, A. Le Pourhiet, A. Rodde, Buffet and buffeting Effect of vortex generators on transonic swept wings, J. Aircr. 53 (6) (2016)
control in transonic flow, Aerosp. Sci. Technol. 9 (7) (2005) 605–616. 1890–1904.
[187] C. Gao, W. Zhang, Z. Ye, Numerical study on closed-loop control of transonic [202] J. Birkemeyer, H. Rosemann, E. Stanewsky, Shock control on a swept wing,
buffet suppression by trailing edge flap, Comput. Fluids 132 (2016) 32–45. Aerosp. Sci. Technol. 4 (3) (2000) 147–156.
[188] J. Liu, Z. Yang, Numerical study on transonic shock oscillation suppression and [203] B. K€onig, M. P€atzold, T. Lutz, E. Kr€amer, H. Rosemann, K. Richter, et al.,
buffet load alleviation for a supercritical airfoil using a microtab, Eng. Appl. Numerical and experimental validation of three-dimensional shock control bumps,
Comput. Fluid Mech. 10 (1) (2016) 529–544. J. Aircr. 46 (2) (2009) 675–682.
[189] R.E. Bartels, J.W. Edwards, Cryogenic Tunnel Pressure Measurements on a [204] E. Stanewsky, J. Delery, J. Fulker, W. Geißler, EUROSHOCK: Drag Reduction by
Supercritical Airfoil for Several Shock Buffet Conditions, NASA-TM-110272, Passive Shock Control; Results of the Project EUROSHOCK, Tech. Rep.; AER 2-
National Aeronautics and Space Administration, 1997. CT92-0049 supported by the European Union 1993–1995, Vieweg, Braunschweig,
[190] T. Barber, J. Mounts, D. McCormick, Boundary layer energization by means of 1997.
optimized vortex generators, in: Proceedings of the 31st AIAA Aerospace Sciences [205] E. Stanewsky, J. Delery, J. Fulker, P. de Matteis, Synopsis of the project
Meeting, Reno, NV, 1993. EUROSHOCK II, in: Drag Reduction by Shock and Boundary Layer Control,
[191] J. Lin, Control of turbulent boundary-layer separation using micro-vortex Springer, 2002, pp. 1–124.
generators, in: Proceedings of the 30th AIAA Fluid Dynamics Conference, Norfolk, [206] S. Raghunathan, J.M. Early, C. Tulita, E. Benard, J. Quest, Periodic transonic flow
VA, 1999. and control, Aeronaut. J. 112 (1127) (2008) 1–16.
[192] J. Dandois, V. Brunet, P. Molton, J.C. Abart, A. Lepage, Buffet control by means of [207] J.P. Eastwood, J.P. Jarrett, Toward designing with three-dimensional bumps for
mechanical and fluidic vortex generators, in: Proceedings of the 5th Flow Control lift/drag improvement and buffet alleviation, AIAA J. 50 (12) (2012) 2882–2898.
Conference, Chicago, IL, 2010. [208] S. Bogdanski, K. Nübler, T. Lutz, E. Kr€amer, Numerical investigation of the
[193] G. Godard, M. Stanislas, Control of a decelerating boundary layer. Part 1: influence of shock control bumps on the buffet characteristics of a transonic
optimization of passive vortex generators, Aerosp. Sci. Technol. 10 (3) (2006) airfoil, in: New Results in Numerical and Experimental Fluid Mechanics IX,
181–191. Springer, 2014, pp. 23–32.
[194] P. Molton, J. Dandois, A. Lepage, V. Brunet, R. Bur, Control of buffet phenomenon [209] T. Streit, K. Horstmann, G. Schrauf, S. Hein, U. Fey, Y. Egami, et al.,
on a transonic swept wing, AIAA J. 51 (4) (2013) 761–772. Complementary numerical and experimental data analysis of the ETW Telfona
[195] J. Dandois, A. Lepage, J.B. Dor, P. Molton, F. Ternoy, A. Geeraert, et al., Open and Pathfinder wing transition tests, in: Proceedings of the 49th AIAA Aerospace
closed-loop control of transonic buffet on 3D turbulent wings using fluidic devices, Sciences Meeting, Orlando, FL, 2011.
Comptes Rendus Mec. 342 (6) (2014) 425–436. [210] S. Bogdanski, P. Gansel, T. Lutz, E. Kr€amer, Impact of 3D shock control bumps on
[196] S. Timme, F. Sartor, Passive control of transonic buffet onset on a half wing-body transonic buffet, in: High Performance Computing in Science and Engineering, vol.
configuration, in: Proceedings of the 16th International Forum on Aeroelasticity 14, Springer, 2015, pp. 447–461.
and Structural Dynamics, St Petersburg, Russia, 2015.

46

You might also like