A Review of SBLI
A Review of SBLI
Keywords: The vast majority of shock wave–boundary-layer interactions in practical applications like supersonic aircraft
Shock–boundary-layer interactions intakes are three dimensional in nature. The complex behaviour of such interactions can generally be
Three dimensionality understood by combining the flow physics of a limited number of canonical cases. The physical understanding
Sidewall effects
of these flow fields developed by numerous investigators over the last half century is reviewed, focusing
Corner effects
predominantly on steady aspects of turbulent, uncontrolled interactions in the transonic and supersonic
Supersonic aerodynamics
Aircraft inlets
regimes, i.e. for Mach number less than 5. Key physical features of the flow fields and recent developments
are described for swept compression corners, various fin interactions, semi-cones, vertical cylinder-induced
interactions, swept oblique shock reflections and flared cylinders. In addition to the canonical geometries,
a different type of three dimensionality concerning sidewall effects in duct flows, like intakes or propulsion
systems, is also reviewed. The underlying mechanisms, centred on pressure waves propagating from the corner
regions, are introduced and the implications for separation unsteadiness and flow control are discussed.
∗ Corresponding author at: School of Engineering and Materials Science, Queen Mary University of London, Mile End Road, London, E1 4NS, United Kingdom.
E-mail addresses: k.sabnis@qmul.ac.uk (K. Sabnis), hb@eng.cam.ac.uk (H. Babinsky).
https://doi.org/10.1016/j.paerosci.2023.100953
Received 10 February 2023; Received in revised form 5 October 2023; Accepted 6 October 2023
Available online 24 October 2023
0376-0421/© 2023 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 1. Examples of canonical three-dimensional SBLIs typically encountered in high-speed aircraft applications.
for the field to progress. Settles and Dolling [9] later provided an This review article briefly summarises the flow physics governing
update to their original review which also included a discussion of com- 2D interactions, as required to introduce the key elements of 3D SBLIs.
putational approaches, unsteady behaviour and flow control. Panaras Following this summary, the key physics of each canonical SBLI are
[10] considered further developments in flow physics understanding, described and recent insight gained alongside directions of current
with non-intrusive experimental methods providing evidence that many research are explained. Finally, the literature on sidewall effects on
of the hypotheses of early researchers were correct. More recently, the SBLIs in channel flows, which is yet to be systematically reviewed, is
additional insights on three-dimensional interactions made possible by surveyed to present the current knowledge of the underlying physical
improvements in simulation capabilities were reviewed by Zheltovodov mechanisms.
and Knight [11]. The continued active interest in 3D interactions is
demonstrated by Gaitonde and Adler [12] who discuss current research
2. A brief summary of two-dimensional SBLI theory
being conducted into the dynamic behaviour of these flow fields.
Over the last two decades, a different type of three dimensionality
has been identified as important and has begun to be widely inves- In an inviscid fluid, a shock wave corresponds to a discontinuous
tigated, namely when an otherwise 2D interaction exists internal to increase in pressure. When the shock impinges on a viscous bound-
a channel. For example, the rectangular engine intakes of supersonic ary layer, however, this pressure rise is significantly smeared. The
aircraft [13], air-breathing missiles [14], and hypersonic propulsion mechanism governing this effect relies on the subsonic channel in
systems [15] have sidewalls. Thus, even nominally 2D interactions are the near-wall portion of the boundary layer. This channel permits
three-dimensional in reality. Similarly, ejectors [16] and wind tunnels upstream travel of pressure information, so that the flow is ‘‘aware’’
used to perform SBLI experiments [17] feature equivalent geometries. of the upcoming pressure rise before encountering the shock itself. The
An analogous flow scenario is encountered when otherwise 2D interac- boundary layer responds to this pressure rise by growing thicker or,
tions are bounded by a sidewall, for example in wing–body junctions. for sufficiently high adverse pressure gradient, by separating. In either
In all these cases the flow field can deviate substantially from two case, the displacement effect of the boundary layer causes streamline
dimensionality. These sidewall effects can be problematic for wind curvature in the external flow. The pressure waves associated with this
tunnel testing [18] and for understanding the key physics from SBLI streamline curvature modify the shock structure, which in turn affects
experiments [19]. Importantly, in practical inlet applications, the same the pressure information transmitted upstream through the subsonic
mechanisms can result in deviations from design behaviour, including channel. The feedback process described above sets up an equilibrium
increased distortion, reduced pressure recovery, the onset of ‘‘buzz’’ flow structure. This type of feedback system underlies every shock
behaviour which precludes unstart, and undesired effects of control wave–boundary-layer interaction. Normal shock interactions, oblique
devices. As a result, there has been considerable effort over the last shock reflections and compression corners do, however, exhibit distinct
two decades to better understand the mechanisms by which sidewalls wave patterns and there are also fundamental differences depending
affect the wider flow field. on whether the boundary layer remains attached or separates. As
2
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 2. Schematic wave patterns for attached and separated flow fields corresponding to the normal shock interaction, oblique shock reflection and compression corner.
the strength of the normal shock increases, the higher pressure rise for any given pressure rise (i.e. shock strength) this pressure gradient is
eventually causes the boundary layer to separate. not fixed but is dependent on the properties of the incoming boundary
The discussion in this article focuses on turbulent shock–boundary- layer. As a result of the sonic line being closer to the wall, a low shape
layer interactions, since they are more prevalent than their lami- factor boundary layer has a smaller subsonic channel to feed pressure
nar counterparts in practical applications which tend to feature high information upstream, which results in a shorter interaction length.
Reynolds numbers. Nevertheless, the flow physics for laminar and The pressure rise is therefore less smeared, and so the boundary layer
turbulent interactions are similar, except that turbulent boundary layers experiences a stronger adverse pressure gradient. The equivalent argu-
are much more resistant to separation, and so the key mechanisms ment also applies in the other direction, i.e. while an increase in shape
introduced can be applied also to laminar cases. Somewhat surprisingly, factor makes a boundary layer more susceptible to separation at a given
the shape factor of the incoming turbulent boundary layer, which adverse pressure gradient, it also increases the upstream influence in an
governs the local momentum, has almost no effect on whether or not SBLI which in turn reduces the adverse pressure gradient imposed on
the boundary layer will separate [2]. the boundary layer. As a result of these two competing effects, the onset
This observation is the result of two competing effects. A low shape of separation is relatively insensitive to boundary-layer shape factor
factor, or higher-momentum boundary layer, does have greater intrinsic and is instead governed primarily by the strength of the shock wave.
resistance to separation for a given adverse pressure gradient. However, Note, however, that the key parameter governing separation is not
3
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 3. Representative streamwise distribution of (a) wall static pressure, 𝑝 (normalised by stagnation pressure, 𝑝0 ) and (b) incompressible shape factor, 𝐻𝑖 , for an attached normal
interaction.
Source: Adapted from Babinsky and Harvey [4]. Reproduced with permission of Cambridge University Press.
Fig. 4. Representative streamwise distribution of (a) wall static pressure, 𝑝 (normalised by stagnation pressure, 𝑝0 ) and (b) incompressible shape factor, 𝐻𝑖 , for a separated normal
interaction.
Source: Adapted from Babinsky and Harvey [4]. Reproduced with permission of Cambridge University Press.
the magnitude of the shock pressure rise itself but rather the adverse thickness, 𝜃𝑖 :
pressure gradient experienced by the boundary layer [20]. Thus, for ( )
𝛿
the same incoming Mach number, if the near-wall portion of the shock 𝛿𝑖∗ ∫0 1 − 𝑢𝑢 𝑑𝑦
𝑒
𝐻𝑖 = = ( ) (1)
wave can be smeared, for example through changes to the surface 𝜃𝑖 𝛿 𝑢
∫0 𝑢
1 − 𝑢𝑢 𝑑𝑦
geometry, then separation can be delayed. 𝑒 𝑒
4
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
5
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
6
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 7. Surface flow topology and schematic of flow in the hinge-normal plane for regimes A – D in swept compression corner interactions. C: compression corner hinge line, U:
upstream influence line, S: separation line, R: reattachment line. Surface flow topologies in regimes B and C based on those proposed by Settles and Teng [45].
7
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
develops into a vortical structure which is directed away from the com-
pression corner tip. The streamsurface originating from the separation
line (the ‘‘separation surface’’) is also entrained into this main vortex.
Meanwhile the flow outside the reattachment surface passes over the
vortical structure [49].
Normal to the hinge line, a two-step rise in pressure is observed. The
initial pressure increase is observed at the separation line, followed by
a plateau region then a further pressure rise at reattachment. Based
on a comparison of several different experiments at Mach 3, Settles
and Dolling [7] state that, for a separated interaction, the ratio of the
plateau pressure to the incoming static pressure is the same as that
measured for a 2D interaction at this Mach number. Nevertheless, it is
worth bearing in mind that wall pressure rises are typically smeared
and so it is difficult to definitively determine the precise plateau
pressure value. As a result, such statements should be interpreted as
order-of-magnitude comparisons rather than the exact values being
Fig. 9. Regimes identified for unswept sharp fin interactions at different fin angles and
identical. In any case, this observation highlights the similarity between
Mach numbers.
swept interactions, when treated in the hinge-normal plane, and their Source: Adapted from Zheltovodov [57]. Reproduced with permission from Springer
two-dimensional counterparts. Settles and Dolling [7] conclude that the Nature.
collapse of pressure profiles implies that the separation mechanism is
governed by the same ‘‘free interaction’’ principles proposed by Chap-
man et al. [50] for 2D interactions, whereby the initial separation is correlation with upstream boundary-layer velocity fluctuations (instead
dependent only on incoming rather than downstream conditions. being driven by structures in the separated region) whereas the low-
frequency dynamics are strongly correlated with the upstream velocity.
Recent research.
The reasons for this discrepancy have not been adequately resolved
Vanstone et al. [51] recently conducted a particle image velocimetry
but one possible reason has been suggested by Vanstone et al. [56].
(PIV) investigation on swept compression corners at Mach 2, in order
When PIV data is compared with fast-response PSP, it is found that the
to assess the extent to which the established scaling laws extend to
characteristic dynamical behaviour of the shock foot and the separation
velocimetry measurements. The collected data implies that, whilst the
line are quite different, thus implying that a more complete picture
streamwise velocity component reliably exhibits the expected conical
of the time-varying flow field is required to fully establish the key
symmetry, the spanwise component only reaches its asymptotic state
mechanisms driving unsteadiness.
after considerably greater distances, sometimes outside experimentally-
accessible regions, and thus the inception region for each velocity
3.2. Single unswept sharp fin interactions
component is different. Until the far-field behaviour is achieved for the
spanwise velocity, a conical flow field can appear to exhibit cylindrical Physical description of flow field.
symmetry. Based on these findings, the authors hypothesise that regime If the swept compression corner is steepened and progressively
B, typically thought to correspond to cylindrical symmetry, might in swept back, it tends towards a sharp fin geometry [46]. The unswept
fact be a conically-symmetric flow which has yet to reach its far-field sharp fin interaction (Fig. 1b) therefore combines features of a stream-
behaviour. However, to date, this hypothesis has not been investigated wise corner interaction with a glancing shock interaction, when a swept
in detail. At lower Reynolds numbers, Padmanabhan et al. [52] used shock wave intersects a boundary layer. Analogous to considering the
computations to study the flow similarity properties for Mach 2.3 swept compression corner interaction in the plane normal to the hinge
laminar swept compression corners, including a comparison with the line, the fin interaction is best treated in the plane normal to the
equivalent turbulent case. The laminar interaction was also simulated glancing shock. In this plane the flow is similar to a normal SBLI. The
at Mach 2.64 in a parametric study conducted by Pulimidi and Lu [53]. behaviour is characterised by the corresponding normal shock Mach
Nevertheless, the majority of recent swept compression-corner re- number, 𝑀𝑛 = 𝑀∞ sin 𝛽0 , where 𝑀∞ is the incoming Mach number and
search has focused on the dynamic behaviour of the turbulent inter- 𝛽0 is the angle of the glancing shock [59]. After an inception zone near
action. The differences between the conical and cylindrical regimes the fin leading edge and before the finite fin extent affects the far field,
were studied by Allen et al. [54], who found that the unsteadiness the flow exhibits conical symmetry in both attached and separated flow
shifts to higher frequencies as the sweep angle increased. Large-eddy cases [7,60–62]. As a result, in spherical coordinates from the virtual
simulations performed by Adler and Gaitonde [34] indicate that the conical origin, the separated flow is self-similar and equivalent to a
low-frequency unsteadiness at St𝛿 < 0.01, which is driven by instabili- normal SBLI with an open separation bubble.
ties in the separated region, is suppressed compared to the equivalent As the deflection angle and, thus, 𝑀𝑛 is increased, several dif-
two-dimensional compression corner interaction. This observation is ferent regimes are identified in Fig. 9. The precise division between
consistent with the swept flow field reducing the dynamic correlation the regimes is a matter of interpretation – Korkegi [63] stated that
between separation and reattachment regions. Meanwhile, the high- there were five regimes whilst Zheltovodov [57] identified six different
frequency behaviour for St𝛿 > 0.1, is found to be mainly driven by regimes. One reason for this discrepancy is the difficulty in interpreting
turbulent fluctuations in the incoming boundary layer in accordance surface oil-flow images to identify separation in three-dimensional
with expectations from 2D SBLIs. However, the dynamics at these flow fields [7,64]. Nonetheless, whatever the precise division between
frequencies are modified by coherent structures in the separated region regimes, the flow progresses from attached to separated as the interac-
which, as they are advected by the cross-flow, cause corrugations in tion strength is increased followed by the appearance, disappearance
the separation line. The signature of these structures are also found and reappearance of secondary separation. Fig. 10 presents the flow
in the mid-frequency range (0.01 < St𝛿 < 0.1) although the pre- topology for increasing values of 𝑀𝑛 . In regime A, which represents
cise mechanisms remain yet to be established. Similar structures have attached flow, Fig. 10a shows no convergence line and the deflection of
also been reported in proper orthogonal decomposition of PIV data surface streamlines through the glancing shock appears similar to the
by Vanstone and Clemens [55]. However, somewhat surprisingly, this expected inviscid behaviour. As the interaction strength is increased,
latter investigation observes that the high-frequency content has little a parallel limiting streamline develops in Fig. 10b, which eventually
8
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 10. Surface flow topology and schematic of flow in the plane normal to the glancing shock for regimes A – E in unswept sharp fin interactions. S: separation line, S1: primary
separation line, S2: secondary separation line, R: reattachment line, R1: primary reattachment line, R2 secondary reattachment line. Flow field schematics based on those proposed
by Zheltovodov [31] and Alvi and Settles [58] and surface flow topologies based on those proposed by Zheltovodov [57].
moves to the outside (i.e. upstream) of the shock wave trace (regime The primary separation results in the formation of a jet, marked
B) [57]. Some authors define this transition to define incipient flow in Fig. 10b, which is bound between the outside of the separated
separation [65], which occurs at a pressure rise ratio of approximately region and the slip line from the lambda-shock triple point. Where the
1.5, corresponding to a normal shock Mach number of 𝑀𝑛 = 1.2. This rear lambda-shock leg encounters the constant-pressure surface of the
value is less than the critical Mach number of 1.3 for the 2D case [66]. vortex edge, it is reflected as a series of expansion waves which turn
Therefore, even though Settles and Dolling [7] report that the pressure the jet towards the wall. When these expansion waves reflect from
rise at the separation line is equivalent to the 2D case in separated the constant-pressure slip surface they form a series of compression
flows, these results imply that cross-flow promotes an earlier onset of waves which can coalesce to form shocklets (Fig. 10c). For the strongest
separation for swept flow fields. interactions in Figs. 10d and 10e, the shocklets become stronger and
9
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
the final, terminating shocklets can form a normal shock [58]. When scales with the incoming boundary-layer thickness, which is consistent
the jet impinges on the wall, some portion of the flow is directed with the large-eddy simulations at Mach 2 by Adler and Gaitonde
towards the fin. There is also a region of reversed flow in the near- [75]. This computational investigation showed that, although this is
wall region of the primary separation vortex, which is directed towards an unswept sharp fin, viscous effects in the inception region cause
the separation line. The behaviour of the separated flow field after the effective aerodynamic shape of the fin’s leading edge to become
the onset of primary separation is affected by the properties of this slightly blunted. As a result, the glancing shock locally takes the form
near-wall reversed flow region [67]. In particular, the presence of any of a bow shock, although the effect of this shock structure transfor-
secondary separation is determined by three factors: (1) the adverse mation on the underlying flow appears to be limited to approximately
pressure gradient imposed on this reversed flow region; (2) the state one boundary-layer thickness. The simulations also revealed deviations
(laminar vs. turbulent) of this flow region; and (3) the acceleration of from perfect conical symmetry which are attributed to the fact that the
this flow region up to supersonic speeds [57,68]. incoming boundary layer is spanwise uniform with a constant thickness
For regime B, the adverse pressure gradient imposed on the cross and thus the flow field cannot be exactly conical. Three-dimensional
flow is insufficient to separate the reversed flow region [68], and so velocity measurements of a Mach 2 turbulent sharp fin interaction using
only primary separation is observed in Fig. 10b. Zheltovodov [69] plenoptic PIV by Jones et al. [76] enabled direct visualisation of the
observed that a further increase in interaction strength causes the devel- overall flow topology by resolving fundamental flow features, such as
opment of secondary separation in regime C (Fig. 10c). This behaviour the lambda-shock structure and the slip region. Further insight into
is attributed to the fact that the near-wall reversed flow is laminar, the flow structure was provided by Zhou et al. [77], who employed
and thus detaches from the wall even for moderate adverse pressure a joint experimental–computational approach to study the evolution of
gradients. Regime C is characterised by the appearance of three critical the vortex structures at Mach 2.95 and Mach 4 for a range of fin angles.
lines: a primary convergence line outside the shock wave trace; a Otten and Lu [78] used Reynolds-averaged Navier–Stokes (RANS)
primary divergence line near the fin–plate junction; and a secondary computations at Mach 2.5 to demonstrate that the flow structure closely
convergence line in between the two [57]. It is worth pointing out that, matches observations on delta wings, and that a gap between the
whilst other researchers also observe a distinct flow feature in the same fin and the flat plate (as often necessitated by practical installation
region as the hypothesised secondary separation, there is some question considerations) introduces an additional vortex into the flow. Liu et al.
as to whether this feature does indeed correspond to a separated region [79] developed a theoretical framework within which the relation-
in regime C [58]. Fig. 10d presents Regime D which corresponds to ship between skin-friction line topology and surface pressure can be
the disappearance of the secondary convergence line, except very near used to understand the onset of secondary separation for unswept
the fin’s leading edge, and so only the primary separation remains. fin interactions. Sebastian and Lu [80] used computations to extend
This is attributed by Zheltovodov [69] to the near-wall reversed flow flow field understanding to laminar interactions, which exhibit the
becoming transitional or turbulent due to the increased streamline same flow structure as their turbulent counterparts with a collapse
length of the bigger separation region. As a result, the cross-flow is in pressure profile when normalised by upstream-influence angle and
able to withstand the imposed adverse pressure gradient and remains plateau pressure.
attached. Note that regime D also corresponds roughly to the point at The surface pressure field of the interaction was surveyed using
which the jet becomes supersonic [68]. Regime E describes the very pressure-sensitive paint by Mears et al. [81] for a sharp fin at 15
strongest interactions, shown in Fig. 10e, where a terminating normal degrees in a Mach 2 flow. A particular focus of the investigation was the
shock forms in the supersonic impinging jet. At this point, it is observed unsteady behaviour of the interaction, which is found to be driven both
that the near-wall reversed flow also accelerates to become supersonic. by the SBLI in the conically-symmetric region but also by flow features
There is a normal shock wave in this near-wall region, whose pressure advecting from the boundary-layer separation at the fin leading edge.
rise is strong enough to separate the flow locally, resulting in the reap- These findings were consistent with the large-eddy simulations of Adler
pearance of secondary separation [70]. As a consequence, secondary and Gaitonde [34], which revealed the same broad conclusions on
convergence and secondary divergence lines are observed very near the unsteadiness as for the swept compression corner interactions, namely
fin [57]. that low-frequency unsteadiness is suppressed while mid-frequency and
As the shock strength increases and different separation regimes high-frequency dynamics are driven by flow structures convected by
are encountered, the corresponding vortex structures change accord- the separation cross-flow.
ingly [71]. The separated free-shear layer rolls into a tight vortex. As for
A further topic of current research concerns the control of unswept
the swept compression corner, flow outside the reattachment surface
sharp fin SBLIs. In order to investigate the key communication path-
continues downstream whilst the flow inside this surface is entrained
ways, flow perturbations at 5.4 kHz were introduced by a microjet
into the vortex [72]. Kubota and Stollery [60] has shown that the flow
array in different regions by Mears et al. [82]. It was found that per-
on the fin itself can also separate, resulting in an additional vortical
turbations introduced upstream of the inception region have a strong
structure on this surface. It is also worth noting that these vortices
effect on the separation bubble. Meanwhile, forcing near the conical
can be smeared and, rather than detecting individual, fully-developed
flow region does not significantly affect the surface pressure but does
vortices, it is more common to observe a supersonic flow along the
cause the mean location of inviscid and separation shocks to move
shock direction with a slow rotational component [73]. Similar to the
upstream. In a separate study by Deshpande and Poggie [83], RANS
swept compression corner, the pressure profile in the plane normal to
computations were used to investigate control of the interaction using
the glancing shock trace corresponds to the 2D equivalent. For the
plasma actuators. This method of control produces a horseshoe vortex,
attached case, a single pressure rise is observed whilst for separated
so that the external flow sees an effective fillet, which makes the
regimes there is a two-step pressure increase, with the plateau pressure
adverse pressure gradient imposed on the flow less severe. As a result,
ratio similar to 2D interactions [48]. It is worth noting, however, that
the separation shock is observed to weaken and the pressure rises
there are some differences such as a pronounced pressure peak (to
over a greater distance [83]. Further RANS simulations at Mach 3
above the inviscid value) at the reattachment points [57].
conducted by Yang et al. [84] at Mach 2.95 studied the effects of the
Recent research. location of surface arc plasma actuation on flow control effectiveness.
The fundamental flow structure of unswept sharp fin interactions Mechanical vortex generators installed in a co-rotating configuration
continues to be actively investigated. A recent experimental investiga- were shown to reduce the strength of the separation shock in an exper-
tion by Baldwin et al. [74] showed that, within the turbulent regime, imental investigation by Verma and Chidambaranathan [85], although
an increased Reynolds number enhances the flow’s resistance to separa- a detrimental strengthening of the separation shock was observed when
tion. The same study also reported that the size of the inception region a counter-rotating set up was used instead.
10
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 11. Schematic flow field at the shock reflection point in a double fin interaction once a Mach stem develops.
11
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 13. Flow field along the axis of a blunt fin interaction at Mach 1.97.
Source: Adapted from Barberis and Molton [107]. Reprinted by permission
of the American Institute of Aeronautics and Astronautics, Inc.
lower. Finally, the position of the initial pressure increase is shifted 3.4. Other interactions being actively studied
to lower angular values, closer to the fin [104]. Nevertheless, when
considered as a function of the Mach number normal to the shock wave, Vertical cylinder-induced interactions.
the plateau pressures for different interaction strengths are the same The flow around cylinders installed on flat plates (Fig. 1g) originally
for swept and unswept fin interactions, as are the critical normal Mach formed the basis of understanding the blunt fin interaction [109] and
numbers for primary and secondary separation [105]. there has been renewed interest in this interaction. Lindörfer et al.
[114] used experiments alongside RANS computations to ascertain the
Perhaps due to the similarities with the unswept fin interactions,
effect of interaction strength (parameterised by the ratio between the
relatively little fundamental research has been performed on supersonic
incoming boundary layer thickness and cylinder diameter) for a Mach
swept fin interactions in recent years. An exception is the recent
2 interaction illustrated in Fig. 14. Once the boundary-layer thickness
numerical study by Otten and Lu [106], who investigated the effects ahead of the cylinder exceeds the cylinder diameter, the curvature of
of a gap between the fin and the wall onto which it is installed, as the shock wave increases significantly, resulting in a greater upstream
commonly occurs in practical applications. These authors found that influence and thus a weakened interaction. The influence of Reynolds
leakage of flow through the tip further weakens the curved shock wave number has also been investigated at Mach 1.8 by Combs et al. [116].
but, when considered in the shock-normal frame, the existing scaling As the Reynolds number is reduced such that the incoming boundary
laws and expected flow structures are observed. layer goes from turbulent to transitional, the mean flow field structure
remains unchanged but a highly intermittent upstream influence wave
Blunt fin interactions. upstream of the leading lambda-shock leg becomes more prominent.
Another canonical modification to the sharp, unswept fin involves In addition, transitional interactions tend to feature a more unsteady
considering a blunt leading edge (Fig. 1e). Such a geometry introduces leading lambda-shock leg than the turbulent counterparts, an observa-
an additional length scale to the geometry, namely the nominal radius tion also made by the same researchers at Mach 4.2 [117]. The dynamic
of curvature at the leading edge [107]. Ahead of the fin, Fig. 13 shows behaviour of turbulent interactions has also been studied for cylindrical
a strong normal shock which always separates the incoming boundary and cuboidal protuberances by Ramachandra et al. [118], who found
layer, even for small fin angles [108]. The flow field in Fig. 13 closely that the frequency of shock oscillation is higher immediately ahead
resembles the interaction for a vertical cylinder installed on a flat of the obstacle than at other spanwise locations. In a more applied
plate [109], as the separated region forms an unsteady vortical struc- geometry, Nimura et al. [119] considered the effect of SBLIs induced by
ture which advects downstream and away from the fin in a horseshoe protuberances with different sizes on the overall force experienced by
vortex system [7]. Early RANS simulations of the flow field by Hung a cone–cylinder geometry which is representative of a slender-bodied
space launch vehicle.
and Buning [110] revealed that the size of this horseshoe vortex and,
therefore, the spatial extent of the interaction is only weakly dependent Swept oblique shock reflections.
on the incoming boundary-layer thickness or the Reynolds number. A theoretical inviscid analysis of swept oblique shock reflections
Sufficiently far from the fin centreline, the leading-edge bluntness no in supersonic flow fields (Fig. 1h) was first conducted by Threadgill
longer affects the flow field and the topology represents that observed and Little [120]. There is a collapse of flow quantities in the plane
for the sharp-fin interaction [111]. normal to the shock sweep angle. Furthermore, the separated flow
Recent research on blunt fin interactions has been performed by field is predicted to display conical symmetry. Evidence for conical
Ngoh and Poggie [112], who used detached-eddy simulations of a Mach behaviour was then found experimentally by Padmanabhan et al. [121]
and Threadgill and Little [115] for a Mach 2.3 interaction (Fig. 15).
3 flow to investigate the mechanisms driving the unsteady behaviour.
These investigations identified quasi-2D regions of flow in suitably se-
The study concluded that the low-frequency motion of the separation
lected planes both upstream and downstream of the inviscid reflection
shock is mainly driven by inherent unsteadiness in the separation
location. The experiments also analysed the low-frequency unsteadiness
region whereas the high-frequency dynamics are mostly related to the
of the interaction, which features large-scale growth and collapse of the
properties of the incoming boundary layer. Castro Maldonado et al. interaction. This dynamic behaviour seems to be caused by unsteadi-
[113] experimentally investigated the effects of fin sweepback at Mach ness mechanisms intrinsic to the separation rather than fluctuations in
4, where the key flow features (vortices, shock–shock interactions and the incoming boundary layer. In particular, these unsteady mechanisms
shear-layer impingement) are captured in the surface heat transfer are related to spanwise-periodic streaks which propagate along the
using temperature-sensitive paint. Consistent with the sharp-fin be- shock foot. The propagation of similar flow features was also identified
haviour, a swept blunt fin is observed to exhibit weakened separation in implicit large-eddy simulations at Mach 2.05 by Lee and Gross [122]
behaviour due to curvature of the shock wave. and in direct numerical simulations performed by Ceci et al. [123]. A
12
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 14. Flow field along the axis of a vertical cylinder interaction at Mach 2.
Source: Adapted from Lindörfer et al. [114]. Reproduced with permission from
Springer Nature.
Fig. 15. Streamwise velocity distribution for a swept oblique shock reflection at Mach 2.3, with flow deflection 12.5 degrees and sweep angle 30 degrees.
Source: Adapted from Threadgill and Little [115].
13
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 17. Comparison between (a) surface flow topology and (b) pressure profile in shock-normal plane for three swept interactions with similar interaction strength at Mach 2.95:
a 30 degree compression corner with 60 degree sweep; a 17.25 degree unswept sharp fin; and a 25 degree semi-cone. C: shock-generator, S1: primary separation line, S2: secondary
separation line, R1: primary reattachment line.
Source: Adapted from Settles and Kimmel [44]. Reprinted by permission of the American Institute of Aeronautics and Astronautics, Inc.
comparison of this flow field versus swept compression corners and proportion of the total cross-sectional area and, for these facilities, the
sharp unswept fins was conducted by Settles and Kimmel [44] for sim- impact of flow confinement is generally assumed to be minor.
ilar inviscid shock strengths. Fig. 17a shows an example of the surface The second, dominant influence of rectangular channel geometries
flow topologies for three such interactions at Mach 2.95, which all is the presence of the corner regions, where the sidewalls meet the
exhibit conical symmetry. It is notable that the primary and secondary floor and ceiling at internal, right-angled corners. These regions are
separation lines collapse on top of one another for the three cases. the intersection of the boundary layers on the horizontal surfaces and
The equivalence of semi-cone interactions with swept compression the sidewalls, and therefore contain low-momentum flow [137]. As a
corners and sharp fins is further highlighted by the collapse of static result, these regions are generally more prone to separation than the
pressure ratio profiles for the three flow fields in Fig. 17b. It is therefore flow away from the corners (Fig. 19). This early separation is in turn be-
generally accepted that the semi-cone flow field can be treated in a lieved to influence the overall flow behaviour in the presence of shock
similar manner to the other swept interactions detailed in the current waves [5]. It is worth noting that the corner regions do have a complex
review. structure, driven primarily by local transfers of momentum towards and
away from the walls by a pair of streamwise counter-rotating vortices
4. Sidewall effects in internal duct flows generated by anisotropies in the turbulent stresses [139]. The precise
structure of these vortices and thus momentum contained within the
Another class of three-dimensional SBLIs concerns flows in rectan- corner region can be influenced by the development history of the
gular duct geometries. The importance of this type of three dimen- boundary layers, such as the nozzle geometry installed in the wind
sionality was first identified through attempts to identify the onset tunnel [140]. In order to obtain quantitatively accurate flow predic-
of separation induced by nominally two-dimensional normal shocks. tions in computations, the presence of the vortices and their precise
As previously discussed, the onset of flow separation in these inter- structure needs to be accounted for to ensure that the size of corner
actions is known to depend primarily on the shock strength, with a separation is correctly captured [141,142]. The requirement to improve
critical shock Mach number of roughly 1.30–1.35 [2]. However, Sajben such calculations has encouraged recent research into the structure of
et al. [136] performed a wind tunnel study at Mach 1.34, finding streamwise corner flows, with a key aim of improving relevant com-
that the flow field remained mostly attached in contrast to previous putational methods [143,144]. Nevertheless, the vortices only cause
expectations. A more comprehensive survey by Bruce et al. [135], relatively small changes to the local flow momentum. Therefore, the
which considered a wide range of experimental facilities, revealed key physical mechanisms can be understood by treating the corner
that separation could occur at a range of shock Mach numbers from flows simply as low-momentum regions, which tend to separate earlier
1.29–1.59 (Fig. 18a). The scatter in data cannot be explained by two- than the flow further away from the corners in response to an adverse
dimensional separation theories. Instead, this observation is attributed pressure gradient (Fig. 19). The effect of this early separation on the
to the wind tunnels used in these experiments having a rectangular flow field depends of the nature of the interaction, with the three
cross section, and thus being fundamentally three dimensional [135]. canonical interactions (normal shocks, oblique shock reflections and
The geometry of a rectangular channel differs from the two- compression corners) considered in turn.
dimensional flow field in two main ways. Firstly, the flow does not
extend to infinity but is confined. The confinement imposes straight 4.1. Corner effects in normal shock interactions
streamsurfaces on symmetry planes and therefore enhances any wall-
normal pressure variations [4]. This effect is exaggerated by the fact The corner separation causes a thickening of the boundary layer
that there are boundary layers developing on all four channel walls downstream of the normal SBLI and thus reduces the effective area of
which thicken in response to the shock pressure rise, thereby reduc- the channel. However, this effect generally has only a minor impact
ing the effective area of the channel [137]. In particularly confined on the flow [145]. The dominant corner influence is instead related
channels, this can cause the nature of the interaction to change. For to the early separation of the corner boundary layer, due to its low
example, an oblique shock reflection can transition from a regular momentum, which causes a displacement effect that deflects the exter-
interaction to an irregular interaction with a Mach stem [138]. Simi- nal flow [146]. Burton and Babinsky [145] proposed that the resulting
larly, a normal shock interaction can turn into a shock train. However, streamline curvature is associated with a series of compression waves
most wind tunnels have boundary layers which only occupy a small which propagate into the flow field. These corner waves influence
14
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 18. The onset of separation with incoming Mach number identified for nominally 2D normal SBLIs across a number of experimental studies. Open symbols denote attached
flow and filled symbols correspond to separated flow. Data plotted as a function of (a) incompressible shape factor of the incoming boundary-layer, 𝐻𝑖 , and (b) viscous aspect
ratio, 𝛿 ∗ ∕𝑊 .
Source: Adapted from Bruce et al. [135]. Reproduced with permission of Cambridge University Press.
the pressure profile experienced by the rest of the flow [147]. For which require higher Mach numbers for centreline separation tend to
simplicity, the compression waves are often treated as a single weak exhibit much larger corner separation [137,150]. To directly test this
conical shock wave, as shown schematically in Fig. 19a [145]. The hypothesis, Bruce et al. [135] performed experiments on a Mach 1.4
corner shock merges with the lambda-shock structure near the floor, normal SBLI which, as shown in Fig. 20, exhibits separation only in the
which is associated with the main normal shock. As a result, there is a corners whilst the centre-span flow remains attached. Targeted suction
greater distance between the front and rear legs of the lambda shock reduces the size of the corner separation but, in doing so, the oil-flow
in the region close to the sidewall. The increased interaction length visualisation in Fig. 20 reveals separation of the previously-attached
reduces the adverse pressure gradient imposed on the flow, and thus
centre-span flow.
delays or decreases the observed separation. Further away from the
It is also believed that the interaction of the corner regions with one
sidewall, the corner effects become weaker and so separation is more
likely [145]. another can introduce subtle asymmetries to the flow field which be-
The most two-dimensional flow field across the channel floor is come more pronounced as the corner separations become larger [151].
expected to be on the centre span, furthest away from the sidewalls. Although similar behaviour has been reported in other studies [16,152,
The extent of shock smearing here depends on the relative size of the 153], the hypothesis has not been investigated in detail and the precise
corner separation to the tunnel width — larger corner separations cause mechanisms behind this effect remain yet to be established.
more pronounced pressure smearing and so a stronger shock is required
to separate the flow. Bruce et al. [135] noted that the corner separation 4.2. Corner effects in oblique shock reflections
is likely to scale with the displacement thickness, 𝛿 ∗ , of the incoming
boundary layer and so proposed this quantity as an alternative param-
The oblique shock reflection interaction in rectangular channels also
eter for the strength of corner effects. Normalising the boundary-layer
displacement thickness by the width of the wind tunnel, 𝑊 , defines exhibits significant three dimensionality, as identified by Reda and
the viscous aspect ratio. When accounting for the effect of confinement Murphy [155] through a direct comparison of the separation behaviour
in this way, the onset of separation induced by normal shocks can be with and without the use of fences to remove the effect of the sidewall
better predicted (Fig. 18b) [135]. Therefore a large corner separation boundary layers. Flow visualisations from a number of studies show
should result in a more benign interaction on the centre span, and vice that the footprint of the interaction on the tunnel floor is significantly
versa. Indeed, facilities with relatively early onset of centreline separa- three dimensional, with curved separation and reattachment lines even
tion tend to feature small corner separations [148,149] whereas those up to the tunnel centreline [156,157]. A parametric study by Sathia
15
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 19. Schematics showing the effect of corner separation on internal duct flow fields for (a) normal shock interactions, (b) oblique shock reflection, and (c) compression corner
interactions.
Fig. 20. Surface oil-flow visualisation of a normal SBLI experiment in a rectangular wind tunnel at Mach 1.4, showing (a) the baseline interaction and (b) a case with corner
separation reduced using local suction. Yellow areas denote any central separation and orange regions correspond to corner separation.
Source: Adapted from Bruce et al. [135]. Reproduced with permission of Cambridge University Press
Narayanan and Verma [158] demonstrates that this three dimension- when the corner compression waves, simplified in Fig. 19b as a con-
ality becomes more pronounced as the incident oblique shock wave ical corner shock, encounter the incident shock wave the behaviour
becomes steeper and the interaction strength increases. In addition to is fundamentally different to the normal shock case. The core flow
noting the curvature of the separation and reattachment lines, Zhang downstream of an oblique shock reflection remains supersonic and
et al. [159] linked the behaviour to the effects of the sidewalls, al- so the corner waves are transmitted through the incident–reflection
though no precise physical mechanism was proposed. Garnier [160] shock system in contrast to the merging behaviour observed for nor-
attributed the poor computational predictions of central separation mal shocks. Fig. 21 presents the footprint of the typical flow on the
to an inaccurate simulation of the corner separation, suggesting that floor, with the corner shocks passing through the incident oblique
corner effects also have an influence on the extent of separation on the shock. Babinsky et al. [161] considered the flow field in three separate
tunnel centreline. regions: (I) the corner separation region, which originates furthest
Similar to the normal shock interaction, the pressure waves gener- upstream; (II) the central separation region, which is often almost
ated by the corner separation provides the fundamental link between two-dimensional; (III) a third intermediate region where the corner
the corner regions and the centre-span separation [161,162]. However, waves smear the pressure rise enough to enable an attached flow
16
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
of a laminar Mach 2 flow. The key difference in laminar flow is the lack
of turbulent stress-induced corner vortices, which results in larger cor-
ner separations and enhanced corner effects. Nevertheless, the central
separation behaviour is related to the waves propagating away from
the sidewalls in the manner predicted by Xiang and Babinsky [154].
However, Lusher and Sandham [165] attributed the origin of these
corner waves to be the footprint of the sharp, unswept fin interaction
on the tunnel floor rather than from the displacement effect of corner
separation. Nevertheless, since the corner wave mechanism appears
to reliably predict the centre-span separation behaviour, recent efforts
have analysed the topology of the entire separation bubble away from
Fig. 21. Schematic flow field for an oblique shock reflection in a duct, divided into the tunnel centre by applying similar principles in streamwise strips
regions: (I) corner separation region; (II) central separation region; (III) intermediate
at different spanwise locations. Missing and Babinsky [166] isolated
region.
the impact of different corner pressure waves by inserting quarter-cone
blockages at the floor–sidewall junction, which impose a known wave
pattern on the flow. At any spanwise location, the separation behaviour
channel. The propagation of the corner shocks towards the centre span of the flow field was explained in terms of the pressure profile imposed
are captured in large-eddy simulations at Mach 2.7 by Wang et al. on the boundary layer by the individual corner compression and ex-
[163]. Depending on the extent of corner separation, different flow pansion waves. A more quantitative approach was adopted by Williams
field regimes are identified in Fig. 22. As well as two-dimensional and Babinsky [167], who considered the pressure profile due to corner
behaviour for large aspect ratios (regime A), it is found that as the waves arriving at several streamwise strips across the tunnel span. By
duct becomes narrower, the streamwise extent of the central separation including the effect of corner compression and expansion waves on
initially increases and then stabilises (regime B). Benek et al. [164]
the overall pressure rise, the resulting strip theory is able to predict
performed RANS computations, also for a range of wind tunnel widths,
variations in the streamwise separation length measured by Xiang and
for a wider range of flow conditions. Prior to the corner effects having
Babinsky [154] across a large proportion of the tunnel. The theory
a significant influence, the authors observe a regime where the central
does break down, however, near the sidewalls where there is a direct
separation length is initially unchanged (regime A). As the corner
interaction between the corner and primary separations.
effects become more important, parameterised by thicker incoming
Recent oblique shock reflection studies have also focused on dif-
boundary layers, the central separation length increases (regime B).
ferent ways in which nominally two-dimensional wind tunnel studies
However, on investigating even narrower ducts, a subsequent decrease
might deviate from expectations. For example, oblique shock gener-
in the streamwise extent of the separated region is observed (regime C).
ators often take the form of a wedge with finite length. Grossman
In an experimental campaign at Mach 2.5, Xiang and Babinsky [154]
and Bruce [168] identified that the expansion fan emanating from the
provided insight into the physical mechanism governing the flow field
downstream corner of this wedge can influence separation properties
by artificially varying the location of the onset of corner separation. As
which are typically attributed solely to the oblique shock reflection.
the corner separation point is moved upstream and the corner shock
The same authors also found that a small gap between the tunnel
intersection location shifts accordingly, regimes A, B and C of the
sidewalls and the nominally full-span shock generator (as often de-
central separation behaviour are all identified in Fig. 22.
manded by practical considerations) can have a measurable impact on
Xiang and Babinsky [154] found that the key influence of the corner
the extent and three dimensionality of the floor separation [169]. More
waves on the separation behaviour at the centre of the tunnel is to
recently, Sabnis et al. [170] demonstrated that the nozzle geometry
modify the adverse pressure gradient imposed on the flow at this cen-
installed to generate the supersonic flow can induce secondary flows
tral location. Fig. 23a shows that in regime A, with the smallest corner
in the sidewall boundary layers, thereby modifying the corner regions
separation, the trajectories of the corner shocks are far downstream of
and influencing the extent of the central separation.
the main interaction in the central part of the tunnel. As a result, these
waves have a negligible impact on the central separation in this region
and the flow field here can be considered quasi-two-dimensional. As the 4.3. Corner effects in compression corner interactions
corner separation point moves further upstream, the corner waves shift
accordingly. Regime B is encountered when the corner shocks intersect Sidewall effects on compression corner interactions were first in-
within the interaction. Fig. 23b demonstrates that these corner waves vestigated in an experimental investigation at Mach 2.56 by Batcho
impose an additional pressure rise on the flow before it can reattach and Sullivan [172]. The footprint of the compression corner interaction
and thus extend the central separation. Even more severe corner sep- exhibits roughly straight separation and reattachment lines across much
aration results in regime C, presented in Fig. 23c, where the corner of the span, and there are no channels of attached flow. Neverthe-
shocks originate far enough upstream that they intersect ahead of the less, the flow field deviates from this behaviour towards the sidewalls
main interaction. Here, the corner separation point is so far upstream with distinct corner separations starting further upstream. In order to
that it is necessary to account for the expansion waves generated by mitigate the effects of the sidewall regions on fundamental numerical
the imposed convex streamline curvature as the footprint of corner studies of this interaction, numerical studies focusing on the 2D com-
separation turns back towards the sidewall. In regime A and B, the pression corner behaviour tend to impose periodic boundary conditions
trajectory of the expansion waves is far enough downstream that they at the sidewalls [173]. For the same reason, experimental compression
do not influence the central separation, but for regime C they intersect corner setups typically suspend a compression corner model in the
within the interaction region. In this regime, the small pressure rise wind tunnel freestream [19] or introduce a gap between the wind
from the corner shocks smears the overall pressure increase over a tunnel sidewalls and the ramp, onto which aerodynamic fences are
greater streamwise extent and the expansion waves reduce the pressure installed [32,174,175]. However, it is worth noting that computations
rise at the shock reflection point. These two effects result in a more which capture the presence of these boundary-layer fences tend to
benign interaction with a central separation which can be even shorter exhibit a central separation which is larger than the equivalent simula-
than regime A. tions with periodic sidewall conditions [176,177]. In addition, Dawson
Similar trends of flow topology, including the three regimes, were and Lele [177] identified spanwise inhomogeneities in large-eddy sim-
reported by Lusher and Sandham [165] from high-fidelity computations ulations of set ups with boundary-layer fences, suggesting that these
17
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 22. Effect of the streamwise position of the corner shock crossing point on centre-span separation length in 8 degree oblique shock reflection experiments at Mach 2.5.
Regimes A – C are identified.
Source: Adapted from Xiang and Babinsky [154]. Reproduced with permission of Cambridge University Press.
Fig. 23. Effect of the corner regions on centre-span separation for a Mach 2.5 oblique shock reflection in regimes A – C, with increasing corner separation extent. For each flow
field, a schematic footprint of the flow field on the tunnel floor, a shock structure schematic on the tunnel centre line and oil-flow visualisation on the tunnel floor are presented.
Source: Adapted from Xiang and Babinsky [154]. Reproduced with permission of Cambridge University Press.
18
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 24. Effect of the streamwise position of the corner shock crossing point on centre-span separation length in 20 degree compression ramp experiments at Mach 2.5. Regimes
A – C are identified.
Source: Adapted from Williams and Babinsky [171].
methods are not completely successful at isolating the interaction from away from the sidewall and modify the imposed pressure profile at
sidewall effects. the centre span. The initial increase in central separation is observed
It is therefore worth considering the behaviour of compression when the corner compression waves intersect over the separated region,
corners which cover the entire duct span, which also feature in practical thus increasing the pressure rise locally. Meanwhile, the subsequent
applications like inlets. The key mechanisms are again related to pres- shortening of central separation corresponds to when the expansion
sure waves emanating from the corner regions, as depicted in Fig. 19c. waves impinge on the central separation, thus reducing the pressure
In studies with ramp angles of up to 15 degrees at Mach 2.5, the corner rise [171]. The effects of the corner compression waves were isolated in
separation is observed to be small in spanwise extent and to be initiated a follow-on study by Teng et al. [182], who used quarter-cone obstacles
only slightly upstream of the primary separation [172,178]. Based at the sidewall–floor junction to remove the influence of the expansion
on large-eddy simulations of a 24 degree ramp at Mach 2.25, Bisek waves and impose conical shocks of known strength from the sidewalls
[179] reported that the central separation is independent of any corner into the core flow. The separation is relatively insensitive to the corner
separation in such flow fields. Nevertheless the presence of waves akin shock strength if the wave trajectories lie upstream of the interaction.
to those reported by Xiang and Babinsky [154] for the oblique shock Once the corner shocks intersect with the separation, there is a larger
reflection can be detected in the surface pressure distributions extracted additional pressure rise imposed on the flow and the size of central
from these computations. separation increases in accordance with expectations. Interestingly,
It is therefore unsurprising that larger corner separations have an once the corner shock waves become strong enough, the flow field
effect on both the flow topology [172,180] and the central sepa- deviates from quasi-two dimensional, with significant corrugation in
ration [181]. The study by Poggie and Porter [181] is particularly
the separation and reattachment lines across the tunnel span [182].
insightful as it studies the same Mach number and ramp angle as Bisek
The spanwise variations in these flow fields are reliably predicted by
[179], but with a flow that is heavily confined in the spanwise direc-
the strip theory proposed by Williams and Babinsky [167], already
tion. The resulting corner separations are pronounced, resulting in a
discussed for oblique shock reflections, which focuses on the effect of
narrow central separation (regime C) which is significantly shorter than
the corner waves on the overall shock pressure rise.
the two-dimensional case (regime A). Experiments on a full-span 20
degree compression ramp at Mach 2.5 by Williams and Babinsky [171]
reveal a further regime (B) between these two extremes. As the corner 4.4. Implications of corner effects for unsteadiness
separation point is moved upstream by using rectangular obstacles to
initiate separation at different streamwise locations, Fig. 24 reveals an The three dimensionality of duct flows has implications to the flow
increase in central separation length from the quasi-2D value in regime steadiness in a way which could be ‘‘critically important for practical
A prior to a subsequent decrease. This observation is analogous to the geometries’’ (Clemens and Narayanaswamy [184]). Therefore, although
trend for oblique shock reflections [154], suggesting that a corner wave the current review is focused primarily on steady flow fields, a brief
mechanism underlies this separation behaviour. comment is included on our current understanding of how sidewall ef-
Indeed, the observed variations in the compression ramp flow field fects relate to dynamic flow behaviour. A review of unsteady behaviour
are explained by Williams and Babinsky [171] in terms of the com- in separated SBLIs by Dussauge et al. [185] first initiated discussion into
pression waves generated by the displacement effect of the corner whether sidewall effects may contribute towards flow unsteadiness and
separation and subsequent expansion waves as the footprint of the cor- what the underlying mechanisms might be. The precise mechanisms are
ner separation turns back towards the sidewall. These waves propagate still not fully understood, but a number of authors have described the
19
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 25. Unsteady behaviour from large-eddy simulations of a 9 degree oblique shock reflection at Mach 2.7. (a) Weighted power spectral density (WPSD) of wall pressures across
duct span. (b) Two-point correlation coefficient from near-wall pressure with reference point on centreline. (c) Two-point correlation coefficient from near-wall pressure with
reference point at sidewall.
Source: Adapted from Rabey et al. [183]. Reproduced with permission of Cambridge University Press.
Fig. 26. Unsteady behaviour from experiments of a 24 degree compression corner at Mach 2.25. Two-point correlation coefficient from wall pressure fluctuations with reference
point at: (a) tunnel centreline, (b) left sidewall, and (c) right sidewall.
Source: Adapted from Deshpande and Poggie [180]. Reprinted with permission from the American Physical Society.
dynamic behaviour of oblique shock reflections and compression corner different to the oblique shock reflection. The large-eddy simulations
interactions. on a 24 degree compression corner at Mach 2.5 by Bisek [179] re-
Large-eddy simulations of an oblique shock interaction (Mach 2.3, veal low-frequency oscillations in both the corner separation and the
9.5 degree flow deflection) performed by Garnier [160] identified central separation. Experiments on highly-confined compression ramp
significant low-frequency unsteadiness in the corner flows, but could interactions by Deshpande and Poggie [180] reveal that, whilst the
not establish a statistical link between these regions and the dynamics ‘‘breathing’’ motion of the central separation has negligible impact on
of the central separation. To investigate this topic further, Rabey et al. the sidewall behaviour (Fig. 26a), the domain of influence of the corner
[183] conducted a joint experimental–computational study of oblique separations extends to 35% of the span which is within the bounds
shock reflections with similar strength in two different aspect ratio of the central separation (Figs. 26b and 26c). Deshpande and Poggie
ducts. The time-varying behaviour across the channel span is presented [186] also conducted dynamic mode decomposition of the flow field,
for one of these cases in Fig. 25. Whilst no direct dynamical link could which suggests that the link between the corner regions and the central
be established between the central and corner separations, the authors separation may be related to an asymmetric alternating expansion and
attributed differences in central separation frequency spectra between contraction of the sidewall separated regions. It seems as though this
the two test cases to duct aspect ratio, implying that sidewall effects behaviour is not limited to severely-confined flows, however, since
might play a role. Indeed, the wall distribution of weighted power a similar asymmetric mode has also been identified using dynamic
spectral density reveals that regions of high low-frequency power in pressure-sensitive paint for a 24-degree ramp at Mach 2.5 in a less
the central separation coincide with locations where the separation confined duct flow by Liu et al. [187].
line is locally displaced upstream. One explanation for this observation Nevertheless, although it was not possible to simultaneously mea-
could be that the mechanism by which the corner regions influence the sure unsteady pressure information in the corners and at the centre span
centre-span dynamics is related to the time-averaged shape and struc- of the wind tunnel, the experiments of Funderburk and Narayanaswamy
ture of the central separation being modified by the aforementioned [178] concluded that there is no dynamical coupling between the
corner waves [154]. two regions due to fundamental differences in the cross-coherence
Compression corner interactions are also observed to exhibit sig- characteristics of the power spectra. Crucially, the corner region dy-
nificant sidewall effects on the central separation unsteadiness [181]. namics show a significant contribution from fluctuations the incoming
However, since the compression corner does not feature attached chan- boundary layer whilst the central interaction responds mainly to fluc-
nels of flow between the corner separation and central separation, the tuations in the downstream separated flow, suggesting an unsteadiness
communication mechanisms between the two regions might be quite mechanism mainly driven by separation bubble instabilities. Williams
20
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
and Babinsky [188] also could not identify a clear correlation between resulted in reduced corner separation. The response of the separation to
corner shock and separation shock unsteadiness for a 20 degree ramp at the control devices is consistent with expected regions of upwash and
Mach 2.5, implying no dynamic coupling between the two regions. The downwash from the vortex generators modifying the local momentum
authors do propose a possible mechanism driving the behaviour, based and thus the susceptibility to separation. Vortex generators placed too
on the observation of Souverein et al. [189] that the power of low- far away from the corners have no effect and those located such that
frequency unsteadiness is primarily influenced by the time-averaged an upwash region coincides with the corner causes the separation to be
size of the separation bubble. Since the streamwise extent of the central aggravated. Therefore, careful positioning of the vortex generators (and
separation is determined by the waves generated due to the displace- knowledge of the subsequent vortex trajectory) is required to ensure
ment effect of corner separation [171], a direct link is established that they produce downwash at the corner boundary layers. Another
between the corner regions and the low-frequency unsteadiness of the possible reason for the inconsistency in vortex-generator effectiveness
central separation without the need for dynamic coupling. between different studies might be that incorrectly configured devices
could disrupt the pre-existing stress-induced corner vortex system [139]
4.5. Implications of corner effects for flow control which can be highly effective at enhancing momentum of the corner
flow. A number of different vortex topologies have been measured in
The close coupling between the corner regions and the flow field corner flows [140], so prior knowledge of the precise vortex system
elsewhere in a duct has important implications for controlling shock- in a given system is necessary to design suitable vortex-generator
induced separation in these flow fields, which is particularly relevant configurations.
for supersonic aircraft inlets, as reported in a recent review by Huang However, due to the corner wave mechanism affecting the pressure
et al. [190]. The flow inside an inlet typically features oblique shock profile away from the sidewalls, control methods successfully applied to
reflections and normal shock interactions, which have both been stud- the corner separation can impact the wider flow field, usually in an un-
ied in inlet-relevant wind tunnel setups with rectangular cross-section. desired manner. For example, Fig. 20 demonstrates how suction applied
Since the low-momentum corner boundary layers separate furthest near the sidewalls is effective in mitigating corner separation but causes
upstream, significant effort has been devoted to reducing separation the appearance of an additional separated region at the centreline.
in these regions. Such research is further motivated by observations The use of magnetically-accelerated surface discharge across the duct
that flow control techniques successfully applied to the central region span is numerically predicted by Atkinson et al. [153] to suppress the
of a rectangular duct tend to exacerbate the separation in the corner corner separation for a 14 degree oblique shock reflection at Mach 2.6.
regions [152]. However, this flow control method simultaneously enhances separation
Titchener and Babinsky [191] found that suction applied to the at the centreline. In a normal shock interaction for an inlet-relevant
corner regions of an inlet-relevant normal shock–subsonic diffuser flow field, Titchener and Babinsky [191] reported that suction applied
setup significantly reduced the corner separations, as demonstrated by to the corner regions results in extensive flow separation on the floor
the comparison between Figs. 27a and 27b. In a very different flow of the diffuser away from the sidewalls (Fig. 27b). The corner fillets
field, Xiang and Babinsky [192] also reported some mild success in mit- effectively employed by Hirt [195] in the corner regions also result in
igating corner separations for a Mach 2.5 oblique shock reflection when an increased boundary-layer thickness at the centreline. Interestingly,
suction slots are installed in the corner regions. However, this study the reduction in corner separation by Xiang and Babinsky [192] for
also suggests that micro-vortex generators installed in the corners, both an oblique shock reflection does not modify the streamwise extent of
on their own and in combination with targeted suction, can be more the central separated region (although it did make the flow field more
successful in reducing corner separation than just suction by itself. two dimensional), perhaps because the studied interaction is already
This apparent success, combined with the substantial size and weight in the quasi-two-dimensional regime where corner effects do not play a
of the apparatus required for targeted bleed in practical applications, role. In the survey of vortex generator configurations to control normal
is reflective of considerable interest in passive flow control methods, shock interactions, Koike and Babinsky [196] note that any set up
including geometry modifications. Babinsky et al. [193] reported an which reduces separation in the corner region also initiates a new
improvement in corner separation for a normal shock–subsonic diffuser separated region on either the bottom wall or the sidewall. However,
setup at Mach 1.3 and 1.5 when corner fences were used. Also con- these particular observations may not be related to waves and could
sidering modifications to the right-angled corner geometry, Baruzzini instead be attributed to the fact that a reduction in corner separation
et al. [194] revealed that chamfering the corner on the scale of the only occurs when there is local downwash in this region, so there
incoming boundary layer thickness is as successful as targeted surface must be upwash elsewhere which makes the local flow more prone to
bleed in reducing corner separation from a Mach 2.5 oblique shock separation.
reflection. Similarly, a study by Hirt [195] showed that applying fillets The close link between different duct regions suggests it is necessary
to the streamwise corners in a Mach 2 oblique shock reflection delays to consider the wider flow field rather than assessing the success of
separation in these corner regions. To date, there are no reports of other a control technique based on analysis of the local effect. Indeed, a
active control methods, such as air-jet vortex generators or plasma number of reportedly successful flow control strategies involve two
actuators, applied specifically to corner separations but since their techniques used in combination to tackle both the corner separation
control mechanism is also centred around manipulation of the incoming and central separation. In an inlet-relevant investigation, Babinsky
boundary layer, similar behaviour is expected to also apply in these et al. [193] found that normal shock-induced separation was prevented
cases. at Mach 1.3 and reduced at Mach 1.5 when micro-vortex genera-
Although vortex generators have been used successfully in other tors at the centreline were used in combination with corner fences
areas including shock–boundary-layer interaction, their effectiveness towards the sidewall. In a similar flow field featuring a Mach 1.4
in reducing corner separations has been inconsistent. In their oblique normal shock interaction, which is presented in Fig. 27, Titchener and
shock reflection study, Baruzzini et al. [194] stated that vane-type Babinsky [191] found that suction applied in the corners combined
micro-vortex generators marginally reduced the corner separations in with vortex generators away from the sidewalls constituted the most
the best-case scenario and, in fact, often had a detrimental effect. effective separation control across the entire tunnel span. As well as
To better understand the flow behaviour, Koike and Babinsky [196] the close interactions already detailed between the corner regions and
studied a large number of vortex-generator configurations in a Mach the central section of a given wall, there are also indications that flow
1.4 normal shock interaction. The authors identified a number of control applied to one wall can influence separation on the other duct
different regimes, depending on the orientation of the vortex generator walls. In simulations of a Mach 2 isolator with the normal shock train
and its distance from the floor–sidewall junction, only some of which producing considerable separation on the top wall, He et al. [197]
21
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Fig. 27. Surface oil-flow visualisation of a normal SBLI experiment, including a subsequent diffusing section, in a rectangular wind tunnel at Mach 1.4, showing (a) the baseline
uncontrolled interaction, (b) suction applied to the corner regions, (c) a vortex generator array installed on the centreline, (d) combined control with corner suction alongside
vortex generators. Yellow areas denote regions of separation.
Source: Adapted from Titchener and Babinsky [191].
attempted to control the flow using suction slots installed in various for a single wall, these studies suggest that, in fact, the communication
configurations. The only arrangement to effectively reduce this corner between different regions demand a more global consideration. This
separation involved suction applied to the top wall near the sidewalls. perspective is especially salient for normal shock interactions, where
However, this control method resulted in significant corner separation the close coupling between central and corner regions is combined with
being introduced instead on the bottom wall. Similarly, Vaisakh and an independent effect, namely that locally reducing flow separation
Muruganandam [198] considered controlling a Mach 1.5 normal shock- in one area changes the post-interaction boundary layer thickness and
induced separation by installing micro-vortex generators on different thus the effective downstream channel size, which in turn affects the
walls of the duct. In general, it was found that applying the control overall post-shock pressure and the nature of the interaction elsewhere.
devices to one wall improved the separation on that wall but enhanced Therefore, all three canonical interactions demonstrate a close re-
the separation on the other walls. It is interesting to note, however, lationship of boundary-layer separation in different flow regions. It
that micro-vortex generators applied to the bottom wall and sidewalls is therefore necessary to consider the entire duct cross-section when
had a positive effect in these regions without any obvious deterioration assessing the impact of applied flow control techniques. For example,
of the separation on the top wall compared to the uncontrolled case. the recent assessments of air-jet vortex generators ahead of a full-
Whilst it is still very common to consider flow control in wind tunnel span 24 degree compression corner in experiments at Mach 2.3 [199]
studies and inlet flows as a local problem that is commonly investigated and computations at Mach 2.5 [200] consider the entire tunnel span.
22
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
Nevertheless, some recent appraisals of the effectiveness of flow control impacting the pressure recovery. It is therefore necessary to simulta-
on rectangular-cross sectioned flow fields for inlet applications consider neously assess both measures of control effectiveness and, importantly,
only the effects of the technique on the centreline rather than over the to bear in mind the requirements of the particular application under
wider region. Examples of such approaches include recent studies in consideration.
micro-jet actuators [201], plasma actuators [202], mechanical vortex It is important to note that wind tunnels used to conduct studies
generators [203] and boundary-layer bleed [204]. These analyses are on shock–boundary-layer interaction and, indeed, many practical ap-
therefore severely limited in their ability to reliably assess the flow plications comprise of nominally-2D SBLIs within confined duct flows.
control effectiveness across the entire flow cross section for practical These flow fields feature boundary layers on the sidewalls as well as the
applications. upper and lower walls. Over the last twenty years, the corner regions
where the sidewalls meet the horizontal walls have been shown to
5. Concluding remarks significantly impact the wider flow field. The underlying mechanism is
caused by low-momentum corner regions tending to separate further
Summary of review. upstream than the flow elsewhere. The displacement effect of the
The relevance of canonical 3D SBLIs to practical applications com- corner separations generate compression and expansion waves which
bined with the inherent complexity of their underlying flow physics propagate away from the corners and modify the pressure gradient
means that these flow fields are still being actively researched more the flow is subjected to, thus affecting the separated flow topology
than sixty years after the first investigations. A review of the phys- for normal shocks, oblique shock reflections and compression corners.
ical flow fields for swept compression corners, fin interactions and These sidewall effects become more extreme for more confined tunnels
semi-cones reveals that, outside the inception region near the leading where boundary layers occupy a larger proportion of the total cross-
edge of the shock generator, semi-infinite interactions tend to display sectional area. As long as the trajectories of the corner waves are
cylindrical symmetry or conical symmetry for weak and strong inter- known, the effect of the corner regions on the flow across the tunnel
actions, respectively. It is often helpful to consider such interactions span can be reliably predicted by considering the pressure profile
in a frame normal to the inviscid shock wave, where the flow struc- imposed at each given spanwise location.
ture can resemble the corresponding 2D interaction. There is also a The corner regions can also influence the dynamic behaviour away
collapse of scaled pressure profiles in this frame, suggesting that these from the sidewalls by modifying the time-averaged central separation,
canonical swept interactions exhibit similar separation flow physics to even though there is broad agreement that there is usually no dy-
each other and to the corresponding 2D cases. As the shock strength is namic coupling between the two regions. The one exception where
increased, a number of regimes are often identified which can progress a direct unsteady coupling has been observed in both high-fidelity
from attached to separated, with subsequent secondary separation. computations and in experiment is the compression corner interac-
This secondary separation region can exhibit quite complex behaviour tion, where asymmetric modes seem to be driving this dynamic link.
including supersonic flow and further shock waves within the primary Establishing the conditions under which different types of dynamic
separation. These separated regions correspond to vortical structures behaviour occur is necessary to develop a complete understanding
which are generally advected downstream and can influence other of unsteady behaviour for practical applications. Various flow control
methods have been successfully applied to the corner regions in order
areas of the flow field.
to reduce the extent of corner separation. Effective techniques include
Recent research has provided new perspectives on these canonical
geometric modifications, targeted surface bleed and micro-vortex gen-
interactions. (1) Experimental studies indicate that the observed cylin-
erators, although a careful consideration of device configurations is
drical regime in all canonical swept interactions may in fact be part of
required. The influence of the corner regions on the wider flow field
a conically-symmetric flow field where, for weak interactions, the far-
means that a reduced corner separation can aggravate the separation
field conical behaviour is only established after significant downstream
away from these regions. Similarly, improving the flow on one wall
distance. The perceived cylindrical symmetry is due to a limited domain
of the duct can worsen the separation on the other walls. This is
of investigation, in which only the inception and mixed regions prior
particularly true for normal SBLIs because of the subsonic nature of
to the far field can be studied. (2) Investigations of dynamic behaviour
the flow downstream, and also because the corner waves tend to be
across the different types of 3D interaction indicate that sweep tends
less steep and thus stronger. It is therefore necessary to consider the
to reduce the amplitude of low-frequency fluctuations (St𝛿 < 0.01),
entire duct cross-section when developing and assessing different flow
which are related to instabilities of the separated region. The high-
control methodologies for these flow fields.
frequency response (St𝛿 > 0.1) is shown to generally be correlated to
turbulence in the incoming boundary layer. However, it is interesting Future prospects.
to note that some experiments on swept compression corners have The recent developments in understanding three-dimensional SBLIs
exhibited the opposite trend, with low-frequency unsteadiness driven naturally lead to several avenues for further investigation, such as
by fluctuations in the incoming boundary layer and high-frequency be- resolving the key mechanisms driving flow unsteadiness in swept inter-
haviour related to coherent structures within the separated region. One actions and developing suitable flow control techniques to mitigate the
possible reason for potential discrepancies between different studies is negative effects of boundary-layer separation. In addition, although the
that they use different flow features as proxies for the overall separa- capabilities of computational methods at predicting three-dimensional
tion unsteadiness. Therefore, a more complete picture of the unsteady SBLIs has not been comprehensively reviewed in the current article,
flow field is required to properly describe the dynamic behaviour and it is worth pointing out that simulations are able to predict these
establish the underlying mechanisms. (3) Another focus of recent 3D complex flow fields with continually-improving accuracy. When RANS
SBLI studies is the control of the separated regions using traditional simulations are unable to predict interactions with corner effects, it
methods such as mechanical vortex generators in addition to novel is usually not because the flow is inherently three dimensional but
techniques like microjet arrays and plasma actuators. There are some because an accurate simulation of the wider flow field requires the
initial indications of the capabilities of different methods to reduce the displacement effect of the corner separation to be correctly captured.
impact of boundary-layer separation but a more systematic survey is Therefore, ensuring reliable prediction of the onset and topology of
required. It is important to note that simply considering one measure the corner separation using low- and medium-fidelity computational
of success is insufficient – a reduction in viscous drag through reducing methods remains a challenge for the future. In this regard, there is
the separation footprint corresponds to a shorter interaction length, considerable current effort to improve the capabilities of RANS sim-
thus strengthening compression waves near the wall and adversely ulations to correctly predict corner separation behaviour by accurately
23
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
reproducing the stress-generated streamwise corner vortices. Neverthe- [21] J. Ackeret, F. Feldmann, N. Rott, Investigations of Compression Shocks and
less, the extent to which such efforts will increase the accuracy of SBLI Boundary Layers in Gases Moving at High Speed, Technical Report, (1113)
National Advisory Committee for Aeronautics, 1947.
simulations for practical applications is yet to be established.
[22] I. Alber, J. Bacon, B. Masson, D. Collins, An experimental investigation
Recent improvements in the accuracy to which three-dimensional of turbulent transonic viscous–inviscid interactions, AIAA J. 11 (5) (1973)
SBLIs can be predicted have both enabled and motivated the study 620–627.
of more complex cases, which more closely resemble practical sce- [23] J. Kooi, Experiment on Transonic Shock-Wave Boundary Layer Interaction,
narios. For example, geometries which are representative of missile Nationaal Lucht-en Ruimtevaartlaboratorium, 1975.
[24] J. Abbiss, L. East, C. Nash, P. Parker, E. Pike, W. Sawyer, A study of the
configurations, slender-body space launch vehicles, and intakes for
interaction of a normal shock wave and a turbulent boundary layer using a
propulsion applications have all been recently investigated in detail. laser anemometer, Technical Report 75141, Royal Aircraft Establishment, 1976.
The latter case corresponds to a double fin interaction with sidewalls, [25] J. Matheis, S. Hickel, On the transition between regular and irregular shock
thus combining both types of three dimensionality discussed in the patterns of shock-wave/boundary-layer interactions, J. Fluid Mech. 776 (2015)
current review. Careful observations of the separated flow fields in 200–234.
[26] J. Délery, Shock wave/turbulent boundary layer interaction and its control,
these more involved cases will enable further insight into the key
Prog. Aerosp. Sci. 22 (4) (1985) 209–280.
physical processes governing three-dimensional SBLIs in the future. [27] G. Settles, S. Bogdonoff, I. Vas, Incipient separation of a supersonic turbulent
boundary layer at high Reynolds numbers, AIAA J. 14 (1) (1976) 50–56.
Declaration of competing interest [28] H. Babinsky, J. Edwards, On the incipient separation of a turbulent hypersonic
boundary layer, Aeronaut. J. 100 (996) (1996) 209–214.
The authors declare that they have no known competing finan- [29] F. Lu, New results on the incipient separation of shock/boundary-layer
interactions, Procedia Eng. 126 (2015) 12–15.
cial interests or personal relationships that could have appeared to
[30] C. Appels, Incipient separation of a compressible turbulent boundary layer,
influence the work reported in this paper. Technical Report 99, Von Karman Institute for Fluid Dynamics, 1974.
[31] A. Zheltovodov, Shock waves/turbulent boundary-layer interactions-
Data availability fundamental studies and applications, in: Fluid Dynamics Conference, AIAA
Paper 1996-1977, 1996.
[32] G. Settles, I. Vas, S. Bogdonoff, Details of a shock-separated turbulent boundary
No data was used for the research described in the article.
layer at a compression corner, AIAA J. 14 (12) (1976) 1709–1715.
[33] G. Settles, T. Fitzpatrick, S. Bogdonoff, Detailed study of attached and separated
References compression corner flowfields in high Reynolds number supersonic flow, AIAA
J. 17 (6) (1979) 579–585.
[1] J. Green, Interactions between shock waves and turbulent boundary layers, [34] M. Adler, D. Gaitonde, Dynamics of strong swept-shock/turbulent-boundary-
Prog. Aerosp. Sci. 11 (1970) 235–340. layer interactions, J. Fluid Mech. 896 (2020).
[2] J. Délery, J. Marvin, E. Reshotko, Shock-Wave Boundary Layer Interactions, [35] J. Nichols, J. Larsson, M. Bernardini, S. Pirozzoli, Stability and modal analysis
Advisory Group for Aerospace Research and Development, Neuilly-sur-Seine, of shock/boundary layer interactions, Theor. Comput. Fluid Dyn. 31 (2017)
France, 1986. 33–50.
[3] P. Viswanath, Shock-wave–turbulent-boundary-layer interaction and its control: [36] J. Dandois, Experimental study of transonic buffet phenomenon on a 3D swept
A survey of recent developments, Sadhana 12 (1) (1988) 45–104. wing, Phys. Fluids 28 (1) (2016) 016101.
[4] H. Babinsky, J. Harvey, Shock Wave–Boundary-Layer Interactions, in: [37] Y. Sugioka, T. Kouchi, S. Koike, Experimental comparison of shock buffet on
Cambridge Aerospace Series, vol. 32, Cambridge University Press, 2011. unswept and 10-deg swept wings, Exp. Fluids 63 (8) (2022) 132.
[5] D. Gaitonde, Progress in shock wave/boundary layer interactions, Prog. Aerosp. [38] J. Délery, Three-Dimensional Separated Flow Topology: Critical Points,
Sci. 72 (2015) 80–99. Separation Lines and Vortical Structures, John Wiley & Sons, 2013.
[6] R. Stalker, Sweepback effects in turbulent boundary-layer shock-wave [39] R. Legendre, Séparation de l’écoulement laminaire tridimensionnel, La
interaction, J. Aerosp. Sci. 27 (5) (1960) 348–356. Recherche Aéronautique (54) (1956) 3–8.
[7] G. Settles, D. Dolling, Swept shock wave/boundary-layer interactions, Tactical [40] L. Squire, The motion of a thin oil sheet under the steady boundary layer on
Missile Aero-dyn. 104 (1986) 297–379. a body, J. Fluid Mech. 11 (2) (1961) 161–179.
[8] A. Smits, S. Bogdonoff, A ‘‘preview’’ of three-dimensional shock-wave/turbulent
[41] M. Tobak, D. Peake, Topology of three-dimensional separated flows, Annu. Rev.
boundary-layer interactions, in: J. Délery (Ed.), Turbulent Shear-Layer/Shock-
Fluid Mech. 14 (1) (1982) 61–85.
Wave Interactions, Springer, 1986, pp. 191–202.
[42] A. Perry, H. Hornung, Some aspects of three-dimensional separation - Vortex
[9] G. Settles, D. Dolling, Swept shock/boundary-layer interactions - tutorial and
skeletons, Z. Flugwiss. Weltraumforsch. 8 (1984) 155–160.
update, in: 28th Aerospace Sciences Meeting, AIAA Paper 1990-0375, 1990.
[43] J. Délery, Physics of vortical flows, J. Aircr. 29 (5) (1992) 856–876.
[10] A. Panaras, Review of the physics of swept-shock/boundary layer interactions,
[44] G. Settles, R. Kimmel, Similarity of quasiconical shock wave/turbulent
Prog. Aerosp. Sci. 32 (2–3) (1996) 173–244.
boundary-layer interactions, AIAA J. 24 (1) (1986) 47–53.
[11] A. Zheltovodov, D. Knight, Ideal-gas shock wave-turbulent boundary-layer inter-
[45] G. Settles, H.-Y. Teng, Cylindrical and conical flow regimes of three-dimensional
actions in supersonic flows and their modeling: Three-dimensional interactions,
shock/boundary-layer interactions, AIAA J. 22 (2) (1984) 194–200.
in: H. Babinsky, J. Harvey (Eds.), Shock Wave-Boundary-Layer Interactions,
[46] G. Settles, On the inception lengths of swept shock-wave/turbulent boundary-
Cambridge University Press, 2011, pp. 202–258.
layer interactions, in: J. Délery (Ed.), Turbulent Shear-Layer/Shock-Wave
[12] D. Gaitonde, M. Adler, Dynamics of three-dimensional shock-wave/boundary-
Interactions, Springer, 1986, pp. 203–213.
layer interactions, Annu. Rev. Fluid Mech. 55 (2023).
[13] Z.-Y. Wang, S. Sun, Y. Zhang, H.-J. Tan, L. Chen, Experimental investigation of [47] G. Settles, J. Perkins, S. Bogdonoff, Investigation of three-dimensional
terminal shock/boundary-layer interaction with and without upstream lateral shock/boundary-layer interactions at swept compression corners, AIAA J. 18
confinement, AIAA J. 61 (1) (2023) 37–47. (7) (1980) 779–785.
[14] K. Boychev, G. Barakos, R. Steijl, S. Shaw, Parametric study of multiple [48] A. Zheltovodov, R. Dvorak, P. Safarik, Shock wave/turbulent boundary layer
shock-wave/turbulent-boundary-layer interactions with a Reynolds stress model, interaction properties at transonic and supersonic speed conditions, Izvestiya
Shock Waves 31 (3) (2021) 255–270. SO AN SSSR, Seriya Tek. Nauk 6 (1990) 31–42.
[15] J. Geerts, K. Yu, Shock train/boundary-layer interaction in rectangular isolators, [49] D. Knight, C. Horstman, S. Bogdonoff, Structure of supersonic turbulent flow
AIAA J. 54 (11) (2016) 3450–3464. past a swept compression corner, AIAA J. 30 (4) (1992) 890–896.
[16] K. Arun, S. Tiwari, A. Mani, Three-dimensional numerical investigations on [50] D. Chapman, D. Kuehn, H. Larson, Investigation of separated flows in supersonic
rectangular cross-section ejector, Int. J. Therm. Sci. 122 (2017) 257–265. and subsonic streams with emphasis on the effect of transition, Technical Report
[17] A. Pope, K. Goin, High-Speed Wind Tunnel Testing, John Wiley & Sons, 1965. NACA-TR-1356, National Aeronautics and Space Administration, 1958.
[18] N. Sudani, M. Sato, H. Kanda, K. Matsuno, Flow visualization studies on [51] L. Vanstone, M. Musta, S. Seckin, N. Clemens, Experimental study of the mean
sidewall effects in two-dimensional transonic airfoil testing, J. Aircr. 31 (6) structure and quasi-conical scaling of a swept-compression-ramp interaction at
(1994) 1233–1239. Mach 2, J. Fluid Mech. 841 (2018) 1–27.
[19] D. Kuehn, Experimental Investigation of the Pressure Rise Required for the [52] S. Padmanabhan, J. Threadgill, J. Little, Flow similarity in swept
Incipient Separation of Turbulent Boundary Layers in Two-Dimensional Su- shock/boundary layer interactions, in: AIAA Scitech 2022 Forum, AIAA Paper
personic Flow, Technical Report, (1-21-59A) National Aeronautics and Space 2022-0942, 2022.
Administration, 1959. [53] R. Pulimidi, F. Lu, On the three-dimensional separation in shock/boundary-
[20] D. Smith, A. Smits, The effects of successive distortions on a turbulent boundary layer interactions at swept corners, in: AIAA Aviation 2019 Forum, AIAA Paper
layer in a supersonic flow, J. Fluid Mech. 351 (1997) 253–288. 2019-3648, 2019.
24
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
[54] D. Allen, K. Langley, J. Schmisseur, P. Kreth, Separation region unsteadiness [84] H. Yang, H. Zong, H. Liang, Y. Wu, C. Zhang, Y. Kong, Y. Li, Swept shock
drivers in swept compression ramps, in: AIAA Scitech 2023 Forum, AIAA Paper wave/boundary layer interaction control based on surface arc plasma, Phys.
2023-2490, 2023. Fluids 34 (8) (2022) 087119.
[55] L. Vanstone, N. Clemens, Proper orthogonal decomposition analysis of swept- [85] S. Verma, M. Chidambaranathan, Swept fin-induced shock/boundary-layer
ramp shock-wave/boundary-layer unsteadiness at Mach 2, AIAA J. 57 (8) separation control using corotating vortex generators, AIAA J. 60 (11) (2022)
(2019) 3395–3409. 6240–6251.
[56] L. Vanstone, T. Goller, L. Mears, N. Clemens, Separated flow unsteadiness in [86] J. Schmisseur, D. Gaitonde, A. Zheltovodov, Exploration of 3-D shock turbulent
a Mach 2 swept compression-ramp interaction using high-speed PSP, in: AIAA boundary layer interactions through combined experimental/computational
Aviation 2019 Forum, AIAA Paper 2019-3647, 2019. analysis, in: 21st Aerodynamic Measurement Technology and Ground Testing
[57] A. Zheltovodov, Regimes and properties of three-dimensional separation flows Conference, AIAA Paper 2000-2378, 2000.
initiated by skewed compression shocks, J. Appl. Mech. Tech. Phys. 23 (3) [87] T. Garrison, G. Settles, Flowfield visualization of crossing shock-
(1982) 413–418. wave/boundary-layer interactions, in: 30th Aerospace Sciences Meeting
[58] F. Alvi, G. Settles, Physical model of the swept shock wave/boundary-layer and Exhibit, AIAA Paper 1992-0750, 1992.
interaction flowfield, AIAA J. 30 (9) (1992) 2252–2258. [88] A. Zheltovodov, A. Maksimov, A. Shevchenko, D. Knight, 3-D separation topol-
[59] R. Korkegi, A simple correlation for incipient-turbulent boundary-layer ogy in asymmetric crossing shock waves and expansion fans/turbulent boundary
separation due to a skewed shock wave, AIAA J. 11 (11) (1973) 1578–1579. layer interaction conditions, Thermophys. Aeromech. 4 (1998) 483–503.
[60] H. Kubota, J. Stollery, An experimental study of the interaction between a [89] T. Garrison, G. Settles, N. Narayanswami, D. Knight, Structure of crossing-
glancing shock wave and a turbulent boundary layer, J. Fluid Mech. 116 (1982) shock-wave/turbulent-boundary-layer interactions, AIAA J. 31 (12) (1993)
431–458. 2204–2211.
[61] G. Settles, F. Lu, Conical similarity of shock/boundary-layer interactions [90] D. Gaitonde, J. Shang, Structure of a turbulent double-fin interaction at Mach
generated by swept and unswept fins, AIAA J. 23 (7) (1985) 1021–1027. 4, AIAA J. 33 (12) (1995) 2250–2258.
[62] S. Garg, G. Settles, Wall pressure fluctuations beneath swept shock [91] P. Batcho, A. Ketchum, S. Bogdonoff, E. Fernando, Preliminary study of the
wave/boundary layer interactions, in: 31st Aerospace Sciences Meeting, AIAA interactions caused by crossing shock waves and a turbulent boundary layer,
Paper 1993-0384, 1993. in: 27th AIAA Aerospace Sciences Meeting, AIAA Paper 1989-0359, 1989.
[63] R. Korkegi, On the structure of three-dimensional shock-induced separated flow [92] K. Poddar, S. Bogdonoff, A study of the unsteadiness of crossing shock
regions, AIAA J. 14 (5) (1976) 597–600. wave turbulent boundary layer interactions, in: 21st Fluid Dynamics, Plasma
[64] S. Bogdonoff, Some observations of three-dimensional shock-wave/turbulent Dynamics and Lasers Conference, AIAA Paper 1990-1456, 1990.
boundary-layer interactions, in: J. Délery (Ed.), Turbulent Shear-Layer/Shock- [93] D. Davis, W. Hingst, Surface and flow field measurements in a symmetric
Wave Interactions, Springer, 1986, pp. 261–272. crossing shock wave/turbulent boundary-layer interaction, in: AIAA Applied
[65] A. Stanbrook, An experimental study of the glancing inter-action between a Aerodynamics Conference, AIAA Paper 1992-2634, 1992.
shock wave and a boundary layer, Technical Report 555, Aeronautical Research [94] D. Knight, T. Garrison, G. Settles, A. Zheltovodov, A. Maksimov, A. Shevchenko,
Council, 1960. S. Vorontsov, Asymmetric crossing-shock-wave/turbulent-boundary-layer inter-
[66] R. Korkegi, Comparison of shock-induced two-and three-dimensional incipient action, AIAA J. 33 (12) (1995) 2241–2249.
turbulent separation, AIAA J. 13 (4) (1975) 534–535. [95] D. Gaitonde, J. Shang, T. Garrison, A. Zheltovodov, A. Maksimov, Three-
[67] F. Lu, G. Settles, Structure of fin-shock/boundary-layer interactions by laser dimensional turbulent interactions caused by asymmetric crossing-shock
light-screen visualization, in: First National Fluid Dynamics Congress, AIAA configurations, AIAA J. 37 (12) (1999) 1602–1608.
Paper 1988-3801, 1988. [96] D. Knight, M. Gnedin, R. Becht, A. Zheltovodov, Numerical simulation of
[68] M. Zubin, N. Ostapenko, Structure of flow in the separation region resulting crossing-shock-wave/turbulent-boundary-layer interaction using a two-equation
from interaction of a normal shock wave with a boundary layer in a corner, model of turbulence, J. Fluid Mech. 409 (2000) 121–147.
Fluid Dyn. 14 (3) (1979) 365–371. [97] D. Gaitonde, M. Visbal, J. Shang, A. Zheltovodov, A. Maksimov, Sidewall
[69] A. Zheltovodov, Properties of two-and three-dimensional separation flows at interaction in an asymmetric simulated scramjet inlet configuration, J. Propuls.
supersonic velocities, Fluid Dyn. 14 (3) (1979) 357–364. Power 17 (3) (2001) 579–584.
[70] F. Alvi, G. Settles, Structure of swept shock wave/boundary-layer interactions [98] Y.-Y. Zhou, Y.-L. Zhao, L. Wang, R. Yang, Y.-X. Zhao, P. Gao, Evolution
using conical shadowgraphy, in: 21st Fluid Dynamics, Plasma Dynamics and characteristics of streamwise vortex of crossing shock wave/turbulent boundary
Lasers Conference, AIAA Paper 1990-1644, 1990. layer interaction, Phys. Fluids (2022).
[71] J. Stollery, Glancing Shock-Boundary Layer Interactions, Technical Report 764, [99] S. Seckin, L. Mears, M. Song, F. Zigunov, P. Sellappan, F. Alvi, Surface
AGARD, 1989. properties of double-fin generated shock-wave/boundary-layer interactions, in:
[72] D. Knight, C. Horstman, B. Shapey, S. Bogdonoff, Structure of supersonic AIAA Scitech 2022 Forum, AIAA Paper 2022-0701, 2022.
turbulent flow past a sharp fin, AIAA J. 25 (10) (1987) 1331–1337. [100] M. Adler, D. Gaitonde, Structure, scale, and dynamics of a double-fin
[73] S. Bogdonoff, The modeling of a three-dimensional shock wave turbulent shock/turbulent-boundary-layer interaction at Mach 4, in: AIAA Scitech 2019
boundary layer interaction, in: 28th Aerospace Sciences Meeting, AIAA Paper Forum, AIAA Paper 2019-0096, 2019.
1990-0766, 1990. [101] M. Schwartz, K. Stamper, R. Bond, J. Schmisseur, Passive flow cntrol on crossing
[74] A. Baldwin, L. Mears, R. Kumar, F. Alvi, Effects of Reynolds number on swept shock-wave/boundary-layer interactions, in: AIAA Aviation 2019 Forum, AIAA
shock-wave/boundary-layer interactions, AIAA J. 59 (10) (2021) 3883–3899. Paper 2019-3595, 2019.
[75] M. Adler, D. Gaitonde, Flow similarity in strong swept-shock/turbulent- [102] L. Wang, Y. Zhao, Q. Wang, Y. Zhao, R. Zhang, L. Ma, Three-dimensional
boundary-layer interactions, AIAA J. 57 (4) (2019) 1579–1593. characteristics of crossing shock wave/turbulent boundary layer interaction in
[76] C. Jones, J. Bolton, C. Clifford, B. Thurow, N. Arora, F. Alvi, Single- a double fin with and without micro-ramp control, AIP Adv. 12 (9) (2022)
camera three-dimensional velocity measurement of a fin-generated shock- 095309.
wave/boundary-layer interaction, AIAA J. 58 (10) (2020) 4438–4450. [103] M. Adler, D. Gaitonde, Influence of separation structure on the dynamics of
[77] Y.-Y. Zhou, Y.-L. Zhao, Y.-X. Zhao, G. He, P.-Y. Gao, Investigation on conical shock/turbulent-boundary-layer interactions, Theor. Comput. Fluid Dyn. 36 (2)
separation vortex generated by swept shock wave/turbulent boundary layer (2022) 303–326.
interaction, Acta Astronaut. 199 (2022) 103–112. [104] A. Zheltovodov, E. Schülein, Problems and capabilities of modeling of turbulent
[78] D. Otten, F. Lu, Flow on leeward side of a sharp fin undergoing swept separation at supersonic speeds conditions, in: Seventh All-Union Congress on
shock/turbulent boundary-layer interaction, in: AIAA Scitech 2023 Forum, AIAA Theoretical and Applied Mechanics, 1991, pp. 153–154.
Paper 2023-1434, 2023. [105] A. Zheltovodov, A. Maksimov, E. Schülein, Development of turbulent separated
[79] T. Liu, D. Salazar, L. Mears, A. Baldwin, Relationship between secondary flows in the vicinity of swept shock waves, in: A. Kharitonov (Ed.), The
separation and surface pressure structure in swept shock-wave/boundary-layer Interactions of Complex 3-D Flows, 1987, pp. 67–91.
interaction, Shock Waves 32 (7) (2022) 665–678. [106] D. Otten, F. Lu, Flow features of swept shock/turbulent boundary-layer inter-
[80] J. Sebastian, F. Lu, Quasi-conical free interaction in fin-induced shock action due to a gap beneath a sharp fin, Aerosp. Sci. Technol. 130 (2022)
wave/laminar boundary-layer interactions, AIAA J. 59 (12) (2021) 4858–4868. 107934.
[81] L. Mears, A. Baldwin, M. Ali, R. Kumar, F. Alvi, Spatially resolved mean and [107] D. Barberis, P. Molton, Shock wave-turbulent boundary layer interaction in a
unsteady surface pressure in swept SBLI using PSP, Exp. Fluids 61 (4) (2020) three dimensional flow, in: 33rd Aerospace Sciences Meeting and Exhibit, AIAA
1–14. Paper 1995-0227, 1995.
[82] L. Mears, N. Arora, F. Alvi, Introducing controlled perturbations in a 3-D swept [108] D. Dolling, S. Bogdonoff, Blunt fin-induced shock wave/turbulent boundary-
shock boundary layer interaction, in: 2018 AIAA Aerospace Sciences Meeting, layer interaction, AIAA J. 20 (12) (1982) 1674–1680.
AIAA Paper 2018-2076, 2018. [109] D. Voitenko, A. Zubkov, Y. Panov, Existence of supersonic zones in
[83] A. Deshpande, J. Poggie, Flow control of swept shock-wave/boundary- three-dimensional separation flows, Fluid Dyn. 2 (1) (1967) 13–16.
layer interaction using plasma actuators, J. Spacecr. Rockets 55 (5) (2018) [110] C.-M. Hung, P. Buning, Simulation of blunt-fin-induced shock-wave and
1198–1207. turbulent boundary-layer interaction, J. Fluid Mech. 154 (1985) 163–185.
25
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
[111] D. Dolling, Comparison of sharp and blunt fin-induced shock wave/turbulent [141] S. Hirt, D. Reich, M. O’Connor, Microramp flow control for oblique shock
boundary-layer interaction, AIAA J. 20 (10) (1982) 1385–1391. interactions: Comparisons of computational and experimental data, in: 5th Flow
[112] H. Ngoh, J. Poggie, Detached eddy simulation of blunt-fin-induced shock- Control Conference, AIAA Paper 2010-4973, 2010.
wave/boundary-layer interaction, AIAA J. 60 (4) (2022) 2097–2114. [142] J. Dandois, Improvement of corner flow prediction using the quadratic
[113] J. Castro Maldonado, J. Threadgill, S. Craig, J. Little, S. Wernz, Flow constitutive relation, AIAA J. 52 (12) (2014) 2795–2806.
structure and heat transfer characterization of a blunt-fin-induced shock- [143] S. Peltier, B. Rice, N. Bisek, C. McKenna, J. Hofferth, Structure of secondary
wave/laminar boundary-layer interaction, in: AIAA Scitech 2021 Forum, AIAA motion in a Mach 2 boundary layer, in: 2018 AIAA Aerospace Sciences Meeting,
Paper 2021-0748, 2021. AIAA Paper 2018-0583, 2018.
[114] S. Lindörfer, C. Combs, P. Kreth, R. Bond, J. Schmisseur, Scaling of cylinder- [144] K. Sabnis, D. Galbraith, H. Babinsky, J. Benek, Experimental validation of the
generated shock-wave/turbulent boundary-layer interactions, Shock Waves 30 quadratic constitutive relation in supersonic streamwise corner flows, in: AIAA
(4) (2020) 395–407. Scitech 2021 Forum, AIAA Paper 2021-1740, 2021.
[115] J. Threadgill, J. Little, Volumetric study of a turbulent boundary layer and [145] D. Burton, H. Babinsky, Corner separation effects for normal shock
swept impinging oblique SBLI at Mach 2.3, Exp. Fluids 63 (9) (2022) 1–20. wave/turbulent boundary layer interactions in rectangular channels, J. Fluid
[116] C. Combs, E. Lash, P. Kreth, J. Schmisseur, Investigating unsteady dynamics of Mech. 707 (2012) 287–306.
cylinder-induced shock-wave/transitional boundary-layer interactions, AIAA J. [146] A. Weber, H.-A. Schreiber, R. Fuchs, W. Steinert, 3-D transonic flow in a
56 (4) (2018) 1588–1599. compressor cascade with shock-induced corner stall, J. Turbomach. 124 (3)
[117] C. Combs, J. Schmisseur, B. Bathel, S. Jones, Unsteady analysis of shock- (2002) 358–366.
wave/boundary-layer interaction experiments at Mach 4.2, AIAA J. 57 (11) [147] T. Handa, M. Masuda, K. Matsuo, Three-dimensional normal shock-
(2019) 4715–4724. wave/boundary-layer interaction in a rectangular duct, AIAA J. 43 (10) (2005)
[118] K. Ramachandra, S. Bhardwaj, S. Rengarajan, Unsteadiness in curved shock- 2182–2187.
induced separation due to protuberances, in: AIAA Scitech 2023 Forum, AIAA [148] T. Bogar, M. Sajben, J. Kroutil, Characteristic frequencies of transonic diffuser
Paper 2023-0242, 2023. flow oscillations, AIAA J. 21 (9) (1983) 1232–1240.
[119] K. Nimura, F. Tsutsui, K. Ktamura, S. Nonaka, Aerodynamic effects of surface [149] P. Doerffer, R. Szwaba, Shock wave-boundary layer interaction control by
protuberance sizes on slender-bodied supersonic vehicle, in: AIAA Scitech 2023 streamwise vortices, in: 11th ICTAM Symposium, 2004.
Forum, AIAA Paper 2023-0241, 2023. [150] W. Sawyer, C. Long, A Study of Normal Shock-Wave Turbulent Boundary-Layer
[120] J. Threadgill, J. Little, An inviscid analysis of swept oblique shock reflections, Interactions at Mach Numbers of 1.3, 1.4 and 1.5, Technical Report 82099,
J. Fluid Mech. 890 (2020). Royal Aircraft Establishment, 1982.
[121] S. Padmanabhan, J. Maldonado, J. Threadgill, J. Little, Experimental study of [151] P. Bruce, H. Babinsky, B. Tartinville, C. Hirsch, Corner effect and asymmetry
swept impinging oblique shock/boundary-layer interactions, AIAA J. 59 (1) in transonic channel flows, AIAA J. 49 (11) (2011) 2382–2392.
(2021) 140–149. [152] N. Titchener, H. Babinsky, Microvortex generators applied to a flowfield
[122] S. Lee, A. Gross, Numerical investigation of sweep effect on turbulent containing a normal shock wave and diffuser, AIAA J. 49 (5) (2011) 1046–1056.
shock-wave boundary-layer interaction, AIAA J. 60 (2) (2022) 712–730. [153] M. Atkinson, J. Poggie, J. Camberos, Control of separated flow in a reflected
[123] A. Ceci, A. Palumbo, J. Larsson, S. Pirozzoli, On low-frequency unsteadiness in shock interaction using a magnetically-accelerated surface discharge, Phys.
swept shock wave–boundary layer interactions, J. Fluid Mech. 956 (2023) R1. Fluids 24 (12) (2012) 126102.
[124] K. Brouwer, R. Perez, T. Beberniss, S. Spottswood, Aeroelastic experiments and [154] X. Xiang, H. Babinsky, Corner effects for oblique shock wave/turbulent bound-
companion computations assessing the impact of impinging shock sweep, in: ary layer interactions in rectangular channels, J. Fluid Mech. 862 (2019)
AIAA Scitech 2023 Forum, AIAA Paper 2023-0945, 2023. 1060–1083.
[125] J. Brown, M. Kussoy, T. Coakley, Turbulent properties of axisymmetric [155] D. Reda, J. Murphy, Sidewall boundary-layer influence on shock wave/turbulent
shock-wave/boundary-layer interaction flows, in: J. Délery (Ed.), Turbulent boundary-layer interactions, AIAA J. 11 (10) (1973) 1367–1368.
Shear-Layer/Shock-Wave Interactions, Springer, 1986, pp. 137–148. [156] P. Bookey, C. Wyckham, A. Smits, Experimental investigations of Mach 3 shock-
[126] M. Garcia, E. Hoffman, E. LaLonde, C. Combs, M. Pohlman, C. Smith, M. wave turbulent boundary layer interactions, in: 35th AIAA Fluid Dynamics
Gragston, J. Schmisseur, Effects of surface roughness on shock-wave/turbulent Conference and Exhibit, AIAA Paper 2005-4899, 2005.
boundary-layer interaction at Mach 4 over a hollow cylinder flare model, Fluids [157] H. Babinsky, Y. Li, C. Pitt Ford, Microramp control of supersonic oblique
7 (9) (2022) 286. shock-wave/boundary-layer interactions, AIAA J. 47 (3) (2009) 668–675.
[127] T. Nilavarasan, G. Joshi, A. Misra, Effect of microramps on flare-induced [158] A. Sathia Narayanan, S. Verma, Experimental investigation of a Mach 4 shock-
shock–boundary-layer interaction, Aeronaut. J. 124 (1271) (2020) 121–149. wave turbulent boundary layer interaction near an expansion corner, in: 53rd
[128] M. Funderburk, V. Narayanaswamy, Investigation of negative surface curvature AIAA Aerospace Sciences Meeting, AIAA Paper 2015-0112, 2015.
effects in axisymmetric shock/boundary-layer interaction, AIAA J. 57 (4) (2019) [159] Y. Zhang, H.-J. Tan, F.-C. Tian, Y. Zhuang, Control of incident shock/boundary-
1594–1607. layer interaction by a two-dimensional bump, AIAA J. 52 (4) (2014)
[129] J. Pickles, B. Mettu, P. Subbareddy, V. Narayanaswamy, On the mean structure 767–776.
of sharp-fin-induced shock wave/turbulent boundary layer interactions over a [160] E. Garnier, Stimulated detached eddy simulation of three-dimensional
cylindrical surface, J. Fluid Mech. 865 (2019) 212–246. shock/boundary layer interaction, Shock Waves 19 (6) (2009) 479–486.
[130] J. Pickles, V. Narayanaswamy, Investigation of surface curvature effects on [161] H. Babinsky, J. Oorebeek, T. Cottingham, Corner effects in reflecting oblique
unseparated fin shock-wave/boundary-layer interactions, AIAA J. 58 (2) (2020) shock-wave/boundary-layer interactions, in: 51st AIAA Aerospace Sciences
770–778. Meeting, AIAA Paper 2013-0859, 2013.
[131] J. Pickles, V. Narayanaswamy, Control of fin shock induced flow separation [162] W. Eagle, J. Driscoll, Shock wave–boundary layer interactions in rectangular
using vortex generators, AIAA J. 58 (11) (2020) 4794–4806. inlets: Three-dimensional separation topology and critical points, J. Fluid Mech.
[132] V. Avduyevskii, V. Gretsov, Investigation of the three-dimensional separa- 756 (2014) 328–353.
tion flow around semicones on a flat plate, Izvestiya Akademii Nauk SSSR, [163] B. Wang, N. Sandham, Z. Hu, W. Liu, Numerical study of oblique shock-
Mekhanika Zhidkosti i Gaza 6 (1970) 112–115. wave/boundary-layer interaction considering sidewall effects, J. Fluid Mech.
[133] J.-H. Liao, X.-Y. Deng, Characteristic flow regimes of swept shock/turbulent 767 (2015) 526–561.
boundary layer interactions, in: 33rd Aerospace Sciences Meeting and Exhibit, [164] J. Benek, C. Suchyta, H. Babinsky, Simulations of incident shock boundary layer
AIAA Paper 1995-0792, 1995. interactions, in: 54th AIAA Aerospace Sciences Meeting, AIAA Paper 2016-0352,
[134] A. Zheltovodov, A. Maksimov, Development of three-dimensional flows at 2016.
conical shock-wave/turbulent boundary layer interaction, Sibirskiy Fiziko- [165] D. Lusher, N. Sandham, The effect of flow confinement on laminar
Technicheskiy Zhurnal 2 (1991) 88–98. shock-wave/boundary-layer interactions, J. Fluid Mech. 897 (2020).
[135] P. Bruce, D. Burton, N. Titchener, H. Babinsky, Corner effect and separation in [166] T. Missing, H. Babinsky, Corner effects on oblique shock wave boundary layer
transonic channel flows, J. Fluid Mech. 679 (2011) 247–262. interactions in rectangular channels, in: AIAA Scitech 2023 Forum, AIAA Paper
[136] M. Sajben, M. Morris, T. Bogar, J. Kroutil, Confined normal-shock/turbulent- 2023-0650, 2023.
boundary-layer interaction followed by an adverse pressure gradient, AIAA J. [167] R. Williams, H. Babinsky, Strip theory approach to corner effects in shock-
29 (12) (1991) 2115–2123. wave boundary layer interactions, in: AIAA Scitech 2023 Forum, AIAA Paper
[137] R. Chriss, W. Hingst, A. Strazisar, T. Keith, An LDA investigation of the normal 2023-0649, 2023.
shock wave boundary layer interactions, La Recherche Aérospatiale (English [168] I. Grossman, P. Bruce, Effect of test article geometry on shock wave-boundary
edition) (2) (1990) 1–15. layer interactions in rectangular intakes, in: 55th AIAA Aerospace Sciences
[138] I. Grossman, P. Bruce, Confinement effects on regular–irregular transition in Meeting, AIAA Paper 2017-0758, 2017.
shock-wave–boundary-layer interactions, J. Fluid Mech. 853 (2018) 171–204. [169] I. Grossman, P. Bruce, Sidewall gap effects on oblique shock-wave/boundary-
[139] F. Gessner, J. Jones, On some aspects of fully-developed turbulent flow in layer interactions, AIAA J. 57 (6) (2019) 2649–2652.
rectangular channels, J. Fluid Mech. 23 (4) (1965) 689–713. [170] K. Sabnis, D. Galbraith, H. Babinsky, J. Benek, Nozzle geometry effects on
[140] K. Sabnis, H. Babinsky, Nozzle geometry effects on corner boundary layers in supersonic wind tunnel studies of shock–boundary-layer interactions, Exp.
supersonic wind tunnels, AIAA J. 57 (8) (2019) 3620–3623. Fluids 63 (12) (2022) 1–16.
26
K. Sabnis and H. Babinsky Progress in Aerospace Sciences 143 (2023) 100953
[171] R. Williams, H. Babinsky, Corner effects for compression corner shock [189] L. Souverein, P. Bakker, P. Dupont, A scaling analysis for turbulent
wave/boundary layer interactions in rectangular channels, in: AIAA Scitech shock-wave/boundary-layer interactions, J. Fluid Mech. 714 (2013) 505–535.
2021 Forum, AIAA Paper 2021-1312, 2021. [190] W. Huang, H. Wu, Y. Yang, L. Yan, S. Li, Recent advances in the shock
[172] P. Batcho, J. Sullivan, The 3-D flowfield in a supersonic shock boundary layer wave/boundary layer interaction and its control in internal and external flows,
corner interaction, in: 26th Aerospace Sciences Meeting, AIAA Paper 1988-0307, Acta Astronaut. 174 (2020) 103–122.
1988. [191] N. Titchener, H. Babinsky, Shock wave/boundary-layer interaction control
[173] K. Porter, J. Poggie, Selective upstream influence on the unsteadiness of a using a combination of vortex generators and bleed, AIAA J. 51 (5) (2013)
separated turbulent compression ramp flow, Phys. Fluids 31 (1) (2019) 016104. 1221–1233.
[174] D. Dolling, C. Or, Unsteadiness of the shock wave structure in attached and [192] X. Xiang, H. Babinsky, An experimental study of corner flow control applied
separated compression ramp flows, Exp. Fluids 3 (1) (1985) 24–32. to an oblique shock-wave/boundary-layer interaction, in: 2018 AIAA Aerospace
[175] M. Ringuette, P. Bookey, C. Wyckham, A. Smits, Experimental study of a Mach Sciences Meeting, AIAA Paper 2018-1532, 2018.
3 compression ramp interaction at Re𝜃 =2400, AIAA J. 47 (2) (2009) 373–385. [193] H. Babinsky, N. Makinson, C. Morgan, Micro-vortex generator flow control for
[176] A. Oliver, R. Lillard, G. Blaisdell, A. Lyrintzis, Effects of three-dimensionality supersonic engine inlets, in: 45th AIAA Aerospace Sciences Meeting and Exhibit,
in turbulent compression ramp shock-boundary layer interaction computations, AIAA Paper 2007-0521, 2007.
in: 46th AIAA Aerospace Sciences Meeting and Exhibit, AIAA Paper 2008-0720, [194] D. Baruzzini, N. Domel, D. Miller, Addressing corner interactions generated by
2008. oblique shock-waves in unswept right-angle corners and implications for high-
[177] D. Dawson, S. Lele, Large eddy simulation of a three-dimensional compression speed inlets, in: 50th AIAA Aerospace Sciences Meeting, AIAA Paper 2012-0275,
ramp shock-turbulent boundary layer interaction, in: 53rd AIAA Aerospace 2012.
Sciences Meeting, AIAA Paper 2015-1518, 2015. [195] S. Hirt, Experimental study of fillets to reduce corner effects in an oblique
[178] M. Funderburk, V. Narayanaswamy, Experimental investigation of primary and shock-wave/boundary-layer interaction, in: 53rd AIAA Aerospace Sciences
corner shock boundary layer interactions at mild back pressure ratios, Phys. Meeting, AIAA Paper 2015-1239, 2015.
Fluids 28 (8) (2016) 086102. [196] S. Koike, H. Babinsky, Vortex generators for corner separation caused by
[179] N. Bisek, Sidewall interaction of a supersonic flow over a compression ramp, shock-wave/boundary-layer interactions, J. Aircr. 56 (1) (2019) 239–249.
in: 53rd AIAA Aerospace Sciences Meeting, AIAA Paper 2015-1976, 2015. [197] Y. He, H. Huang, D. Yu, Investigation of corner separation and suction control
[180] A. Deshpande, J. Poggie, Large-scale unsteadiness in a compression ramp flow in constant area duct, Aerosp. Sci. Technol. 66 (2017) 70–82.
confined by sidewalls, Phys. Rev. Fluids 6 (2) (2021) 024610. [198] S. Vaisakh, T. Muruganandam, Influence of multi-wall separation control on
[181] J. Poggie, K. Porter, Flow structure and unsteadiness in a highly confined normal-shock-induced separation in supersonic duct flows, Proc. Inst. Mech.
shock-wave–boundary-layer interaction, Phys. Rev. Fluids 4 (2) (2019) 024602. Eng. G 233 (9) (2019) 3184–3192.
[182] L. Teng, R. Williams, H. Babinsky, The influence of conical shocks on compres- [199] D. Ramaswamy, A.-M. Schreyer, Control of shock-induced separation of a
sion corner shock wave/boundary layer interactions, in: AIAA Scitech 2022 turbulent boundary layer using air-jet vortex generators, AIAA J. 59 (3) (2021)
Forum, AIAA Paper 2022-0699, 2022. 927–939.
[183] P. Rabey, S. Jammy, P. Bruce, N. Sandham, Two-dimensional unsteadiness map [200] R. Sebastian, A.-M. Schreyer, Control of SWBLI in a 24deg compression ramp
of oblique shock wave/boundary layer interaction with sidewalls, J. Fluid Mech. flow with air-jet vortex-generator, in: AIAA Scitech 2023 Forum, AIAA Paper
871 (2019). 2023-2138, 2023.
[184] N. Clemens, V. Narayanaswamy, Low-frequency unsteadiness of shock [201] M. Chidambaranathan, S. Verma, E. Rathakrishnan, Control of incident shock-
wave/turbulent boundary layer interactions, Annu. Rev. Fluid Mech. 46 (2014) induced boundary-layer separation using steady micro-jet actuators at 𝑀∞ = 3.5,
469–492. Proc. Inst. Mech. Eng. G 233 (4) (2019) 1284–1306.
[185] J.-P. Dussauge, P. Dupont, J.-F. Debiève, Unsteadiness in shock wave boundary [202] X. Ma, J. Fan, Y. Wu, S. Zhu, R. Xue, Flow control effect of pulsed arc discharge
layer interactions with separation, Aerosp. Sci. Technol. 10 (2) (2006) 85–91. plasma actuation on impinging shock wave/boundary layer interaction, Phys.
[186] A. Deshpande, J. Poggie, Dynamic mode decomposition of a highly confined Fluids 35 (3) (2023) 036110.
shock-wave/boundary-layer interaction, in: AIAA Scitech 2021 Forum, AIAA [203] M. Kaushik, Experimental studies on micro-vortex generator controlled
Paper 2021-1097, 2021. shock/boundary-layer interactions in Mach 2.2 intake, Int. J. Aeronaut. Space
[187] X. Liu, L. Zhang, Y. Ji, M. He, Y. Liu, D. Peng, Fast PSP measurement Sci. 20 (3) (2019) 584–595.
of three-dimensional low-frequency unsteadiness in sidewall-confined shock [204] N. Thillai, A. Thakur, K. Srikrishnateja, J. Dharani, Analysis of flow-field in a
wave/turbulent boundary layer interaction, Exp. Therm Fluid Sci. 134 (2022) dual mode ramjet combustor with boundary layer bleed in isolator, Propuls.
110599. Power Res. 10 (1) (2021) 37–47.
[188] R. Williams, H. Babinsky, Corner effects on the unsteady behaviour of com-
pression corner shock wave/boundary layer interactions, in: AIAA Scitech 2022
Forum, AIAA Paper 2022-1181, 2022.
27