A Review of Impinging Jets During Rocket Launching

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Progress in Aerospace Sciences 109 (2019) 100547

Contents lists available at ScienceDirect

Progress in Aerospace Sciences


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

A review of impinging jets during rocket launching T


Chongwen Jiang , Tianyixing Han, Zhenxun Gao, Chun-Hian Lee

National Laboratory for Computational Fluid Dynamics, School of Aeronautic Science and Engineering, Beihang University, 100191, Beijing, China

ARTICLE INFO ABSTRACT

Keywords: During rocket launching, the engine exhaust impinges on the launch structures, such as the launch platform and
Impinging jet the jet deflector, leading to a complex flowfield and generating strong noise. The engine exhaust impingement
Rocket launch induces large aerothermodynamic and acoustic loads on the launch structures and the rocket, the overload of
Flow structure which is a potential risk to the rocket launch. Thus, it is important to predict and reduce the loads. Although the
Noise
jet impingement during rocket launching has been studied extensively, the jet impingement during rocket
Launch pad
Multiphase flow
launching is a complex phenomenon, where multiple mechanisms are coupled. In most researches, several
mechanisms are studied separately or in a partially-coupled way. Still, there are a lot of issues on the jet im-
pingement remaining to be solved, especially when quantitative relations are taken into account.
In this article, several issues involved are reviewed, to present an overview on progresses and perspectives of
the jet impingement research for rocket launching. It starts with a summary of the results accessed for normal
and inclined jets impinging on flat plates, then the interactive mechanisms between impinging jets and the
launch platform or the deflector system, together with relevant factors that affect the flowfield and noise.
Additionally, discussions on the interaction of impinging jets with multiphase flows are also presented.

1. Introduction flow could become separated and generate a recirculation region [24].
Flow structures of the impinging jet, especially the near-wall flow field,
Flows induced by impinging jets are commonly encountered in may influence strongly the aerodynamic loads and the thermal loads
many aerospace applications, such as in the launch site during a rocket [26] on impinged objects. Meanwhile, disturbances in the shear layer
launching [1–3], V/STOL (Vertical and/or Short Take-Off Landing) would induce oscillations of the near-wall flow structures, and thus
aircraft hovering near the ground [4,5], and the extraterrestrial landing influences the acoustic loads [27]. In addition, multiphase flow induced
and launching of spacecraft [6,7]. Characteristics of impinging jets are by water injection [1,20] or solid propellant combustion [28,29] might
different from coflow jet [8–10] and lateral jet [11–15]. During rocket appear in the engine exhaust impingement.
launching, the engine exhaust is usually a supersonic over-expanded jet As shown in Fig. 2, during the rocket launching, the flowfield of
[1–3,16]. engine exhaust impingement behaves differently at varying lift-off
Large rockets are usually launched from launch pads, which include heights, which is defined as the distance between the nozzle exit and
several structures [1,20], such as the service structure, the launch the launch platform. Thus, the flowfield over the rocket changes dy-
platform and the deflector system (deflector, deflector duct/trench, namically during its lift-off stage. In the ignition stage as shown in
duct/trench cover), as shown in Fig. 1. The service structure is near the Fig. 2a, the engine is ignited and works stably, producing engine ex-
rocket, providing supplies (e.g. propellants, electrical power, commu- haust. The engine exhaust flows through the exhaust vent on the launch
nication parts) before launching. The launch platform lies beneath the platform and reaches the deflector duct, producing certain flow struc-
rocket with cut-outs at some specific positions. The engine exhaust tures near the deflector exit. When the rocket starts to ascend (c.f.
passes through the cut-outs, impinges onto the deflector and deflects Fig. 2b and c), the engine exhaust impinges onto the launch platform
away from the rocket. During the launching, exhaust of the rocket en- and produces complex flow structures nearby. During this stage, the
gine possibly impinges on the launch platform, the deflector system or engine exhaust mainly interacts with the launch platform or the de-
the ground [21–23]. Characteristics of the impinging jets are different flector system. As the rocket ascends further (Fig. 2d), the interaction
from the free jets. The impinging jet may induce standoff shocks (or between the engine exhaust and launch structures becomes weaker. The
plate shocks) and wall jets. Under certain conditions, the impinging jet engine exhaust tends to be a free jet. The field measurements [23]


Corresponding author.
E-mail address: cwjiang@buaa.edu.cn (C. Jiang).

https://doi.org/10.1016/j.paerosci.2019.05.007
Received 28 December 2018; Received in revised form 23 May 2019; Accepted 25 May 2019
Available online 31 May 2019
0376-0421/ © 2019 Elsevier Ltd. All rights reserved.
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

Fig. 1. Launch pads. (a) Vostochny Cosmodrome Site 1S, Amur Oblast, Russia. Reproduced from Ref. [17], licensed under Creative Commons Attribution 4.0
International license. (b) Mid-Atlantic Regional Spaceport (MARS) Launch Pad 0A, Virginia, USA. Reproduced from Ref. [18], licensed under the Creative Commons
Attribution 2.0 Generic license. (c) Rocket Lab Launch Complex 1, Mahia, New Zealand. Reproduced by permission from Rocket Lab [19]. (d) Scheme of a typical
launch pad. Reproduced by permission from SAGE Publications [20].

indicated that, during the intermediate stage as shown in Fig. 2b and c, applied to the prediction of the acoustic characteristics of free and
the launch platform and the deflector system are main noise sources at impinging jets [32,33] as well as the acoustic loads during rocket
different lift-off heights. The sources would produce strong acoustic launching [34–41]. However, the source allocation methods require
loads on the launch structures and the rocket, especially the payload sufficient noise measurements to produce a robust prediction on the
and electronic systems. Thus, the interaction with the launch platform spectra at the points far from the measured points.
and the interaction of the engine exhaust with the launch platform and There involves in the jet impingement problem of rocket launching
deflector system are two of the key issues related to the engine exhaust the multi-physicochemical mechanisms coupled with flow-structural
impingement during the rocket launching. Meanwhile, the jet im- interactions. The launch platform has shelter or reflection effects on
pinging on launch structures produces large thermal loads and possible acoustic waves generated from different sources [42,43], but itself may
ablation on the structures [30]. Hence, it is vital for the launch pad become simultaneously a noise source when engine exhaust impinges
design, to fully access the information on the interactive flowfield of on it [21–23]. Similarly, jet deflectors may cause side effects when the
engine exhaust impingement and predict aerothermodynamic and engine exhaust impinges on them. For single-sided deflectors (i.e., de-
acoustic loads accurately. flectors which deflect engine exhaust towards a single direction), de-
As indicated in the NASA Report SP-8072 [31], the acoustic loads pending on the state of the expansion jet and the impingement distance,
generated by the propulsion system are usually predicted by empirical there may cause various types of interactive flow structures [25,44,45],
analyses and sub-scale model experiments, where the information of which would then generate various types of acoustic waves in the
the noise sources attained from experiments is used to allocate flowfield [46]. A water injection system is utilized in some launch pad
equivalent sources along the engine exhaust and then reconstruct the to reduce thermal and acoustic loads induced by the exhaust im-
acoustic field. The source allocation methods have been developed and pingement. During the rocket launching, water jets may interact with

2
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

Fig. 2. Different phases of a rocket launch. The engine is ignited at t = 0 . Reproduced by permission from NASA/Jayanta Panda [23].

engine exhaust, possibly resulting in break-up, atomization and phase constitutes two basic forms, namely, the normal and inclined jets im-
change [47]. As to solid rocket engine, discrete particles are generated pinging on plates resembling the launch platform and the deflector.
in the combustor, break up in the nozzle and interact with the engine Simplified models of these two configurations of impingements have
exhaust. The discrete particles would exchange momentum and energy been extensively studied, including multispecies and multiphase flow
with the engine exhaust, gathering with phase changes during the characteristics [53–55], chemical reactions [30], and the movement of
whole process [28,48]. Chemical reactions in the combustor generate discrete particles [56–59], etc. However, there are more fundamental
high temperature exhaust, and may induce afterburning when the re- issues related to the practical problems remained to be further in-
action products are mixed with surrounding fluid [49]. The above vestigated.
stated phenomena could happen simultaneously and couple with each In this article, several issues involved are reviewed, and the ad-
other. While the afterburning of the engine exhaust imposes physico- vances on the studies of interactive flowfields and noise of impinging
chemical effects on the discrete particles [28], the afterburning itself is jets are discussed. It starts with a summary of the results accessed for
influenced by the water injection [30,50]. All the phenomena have a the two simplified models, i.e. normal and inclined jets impinging on
combined impact on flow characteristics of the engine exhaust as well flat plates in Sec. 2. The interactive mechanisms between impinging jets
as the aerothermodynamic and acoustic loads of the launch pad struc- and the launch platform or the deflector system are then presented in
tures. Sec. 3, together with relevant factors that affect the flowfield and noise.
Field measurements of rocket launching [23] provide the informa- In Sec. 4, discussions are directed toward the interaction of impinging
tion on the interaction of the impinging flow with actual structures of jets with multiphase flows, where special emphases are paid to the
the launch pad under real conditions, including all the physicochemical noise suppression of water injections and the effects of discrete particles
mechanisms. However, it is difficult and expensive to measure the on flow characteristics. Finally, in Sec. 5, an overview on progresses
impinging flowfield of the engine exhaust in details. As a result, field and perspectives of the jet impingement research for rocket launching is
measurements are difficult to repeat and few works have ever been presented.
published. Model experiments with reduced scales are usually em-
ployed to study the phenomena [1–3,21,22,41,42,51]. The reduced
scale models would include typical structures involved in the actual 2. Impinging jet interaction with a plate
rocket launch pad. Influences of the launch platform and the deflector
can be studied at certain controlled conditions. It is difficult, however, As designated to be simplified models for the two configurations of
to preserve the similarity between the scale model experiment and the engine exhaust impingements, the normal and inclined jets impinging
actual launching. Numerical studies, on the other hand, are capable of on plates have been extensively studied [24,44,60]. Studies conducted
investigating the complex interaction of flow impingement onto launch on the former models over the past 60 years can serve as important
pad structures to some extent [43], but their feasibility is constrained references for understanding the latter configurations, including the
by computational resources and algorithmic efficiency [52]. In addi- classification of different types of the impinging jet noise, their re-
tion, the uncertainties imposed by the modelling of physicochemical spective mechanisms of the noise production and propagation [46,61],
processes may further affect the reliability of the numerical approaches. as well as the relation between the impinging noise and the typical flow
The engine exhaust impingement related to the rocket launching structures of impinging jets [24,46,61]. Some typical flow structures of
impinging jets will be reviewed in this section, together with the nature

3
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

Fig. 3. Mean shadowgraphs of under-expanded supersonic impinging jets at varying im-pingement distances. Mj = 1:5 and NPR = 5:0. NPR stands for nozzle
pressure ratio, stagnation pressure at the nozzle inlet/ambient static pressure. Reproduced from Ref. [63] with permission of Cambridge University Press.

of their respective noise production and propagation.

2.1. Flow structure of impinging jets

2.1.1. Flow structure of normal impinging jets


The nearfield flow structures of a jet depend primarily on the op-
erational conditions of the nozzle, which are characterized mainly by
the Mach number of the jet at the nozzle exit (Mj) and the ratio of the
static pressure at the nozzle exit to the ambient static pressure (PR), for
either free jets [62] or impinging jets [24,44]. For impinging jets, the
whole flowfield would also be affected by the impingement distance
(h), i.e., the distance from the nozzle exit to the intersection point of the
jet axis and the impinged surface. In practice, its nondimensional form
is preferred, which is usually normalized by the nozzle exit diameter
Fig. 4. Scheme of a normal impinging jet.
(d). However, under fixed nozzle operational conditions, the first shock
cell is insensitive to the varying impingement distance [26,63,64], as
shown in Fig. 3. It turns out that flowfields of the normal impinging jet
are similar to those in the free jet in a certain region near the nozzle
exit, which is called the free jet region [60]. The region downstream of
the free jet region and close to the plate is called the impingement re-
gion [24,60,65], while the region where the wall jet emerges is called
the wall jet region [24,60,65], (c.f. Fig. 4).
Flow structures in the free jet region resembles a similar feature as
in the free jet under the same nozzle operational conditions, as illu-
strated in Fig. 5. For an under-expanded jet, the flow expands at first
until its pressure balances with the ambient pressure. Then, the flow
becomes over-expanded and induces compression waves, which con-
verge to form the intercepting shocks [62]. For an over-expanded jet, in
other words, the pressure at the nozzle exit is lower than ambient
pressure, oblique shocks would be generated by the compression from Fig. 5. Scheme of a moderately under-expanded free jet.
the surrounding fluid [66]. For an ideally expanded jet, the oblique
shocks become weaker and are not as distinguishable as in the over-
the reflected shocks may resemble an X shape (or a conical shape) if the
expanded case [67,68]. The velocity difference across the pressure-
reflection is regular [24], or a Mach disk if the reflection is a singular
balanced boundary creates a shear layer (or mixing layer), which tends
[62]. Slipstreams would appear at the triple points, where the incident
to grow downstream. At the same time, the incident oblique shocks
shocks, the reflected shocks and the Mach disk intersect. When the jet is
reflect at certain points in the vincinity along the jet axis. Structures of

4
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

to where the first shock cell of the free jet under the same nozzle op-
erational conditions attains the maximum diameter). In type II, the
reflection of the upper intercepting shock by the plate would generate a
reflected shock called the upper plate shock, which interacts with the
lower intercepting shock and generate a normal shock called the lower
plate shock. Two slipstreams would originate from the triple points.
Below the lower slipstream, quasi-periodic shock cells could be ob-
served. Under a higher pressure ratio, a smaller inclined angle or a
smaller impingement distance than type II, a variant of type II, denoted
by type II′, could appear. The flow structures of type II’ are almost the
same to type II, while a stagnation bubble would emerge near the upper
tail shock. With an even smaller h/ h s , the flowfield appears as type III,
where the lower plate shock shrinks and the slipstreams merge into one.
The connection between several flow structures and the pressure dis-
tribution on the plate has been established. Four sources of the pressure
peak could exist, including the stagnation of the main jet, the stagnation
below the upper tail shock, the reattachment of the detached flow (if
Fig. 6. Positions of shock structures at varing impingement distances. Mj = 1.0 , the stagnation bubble exists) and the reflection of the intermediate tail
NPR = 3.0 . Reproduced with permission of ASME, from Ref. [71]; permission shocks downstream of the impingement region. By combining the
conveyed through Copyright Clearance Center, Inc. classification of the flow structures [44] and the pressure peeks, the
inclined impinging jets could be further classified into six types [45].
not extremely under-expanded, quasi-periodic shock cells could appear
repeatedly [62]. The shear layers would merge at the jet axis as the jet 2.2. Noise production of impinging jets
develops downstream, causing nearfield shock structures disappear and
thence the potential core of the jet ends. When the impinged plate is The flowfield of the impinging jet could be divided into the free jet
near the potential core, there would be further interaction between the region, the impingement region and the wall jet region. The flow
impinging jet and the plate [69,70]. structures of the impinging jet in the free jet region show similarity to
In the impingement region, a detached shock or a recirculation re- those of the free jet [60], as well as the acoustic characteristics. In
gion may form under certain jet Mach number, pressure ratio and im- addition to the noise originated from the free jet region, there is also
pingement distance. If the shock cell approaching the plate resembles noise reflected by the plate and noise induced by the interactions
an X shape, the flow downstream of the oblique shocks is supersonic among the free jet region, the impingement region and the wall jet
and a detached shock called a standoff shock usually appears to meet region.
the pressure balance [24]. As illustrated in Fig. 6, by varying the im-
pingement distance and taking the pressure ratio and jet Mach number 2.2.1. Noise of free jets
fixed, the distance between the standoff shock and the plate, as well as Historically, noise of free jets has long been studied [61,77–80].
the locations of the shock cells upstream of the standoff shock, is almost According to the review by Tam [61], except for the perfectly expanded
constant [71]. If the shock cell close to the plate contains a Mach disk, jet, noise of a supersonic free jet comprises three basic components: the
the total pressure loss downstream of the normal shock is higher than turbulent mixing noise, the broadband shock-associated noise and the
the reflected shocks. This leads to the stagnation pressure at the inter- screech tone.
section point of the jet axis with the plate to be lower than the ones in The turbulent mixing noise is generated by the downstream con-
the surrounding, which causes a recirculation region with a wedge or vection of the large turbulent structures [81–83]. The dominant pro-
conical shape to form in front of the plate, as shown in Fig. 4. The flow pagation direction is 45 –60 measured from the downstream direction
outside the recirculation region is compressed and thus induces the of the jet axis, and the dominant Strouhal number (St) is 0.1–0.25. The
attached oblique shock called the annular shock. When the flow is re- turbulent mixing noise is mainly affected by the jet Mach number and
accelerated to be supersonic in the wall jet region, a trains of shock cells the jet temperature. When the jet Mach number increases, the sound
similar to the quasi-periodic structures in a free jet may form along the pressure level (SPL) of the turbulent mixing noise above the back-
wall jet [72]. ground noise increases. As the jet temperature increases, the dominant
band extends towards lower frequency in the spectrum, and thus the
2.1.2. Structure of inclined impinging jets overall sound pressure level (OASPL) increases.
The flow structures of inclined impinging jets resemble those of The broadband shock-associated noise is induced by the interaction
normal impinging jets under a small inclined angle, while the flow between the turbulent structures and the shock structures [82–87]. The
downstream of the impingement region would be more complex as the dominant frequency is higher than that of the turbulent mixing noise.
inclined angle increase [73–75] (c.f. Fig. 7). Except for the inclined When the observation angle increases, both the dominant frequency
angle, the pressure ratio and impingement distance could also affect the and the bandwidth decreases. As the jet temperature increases, the
flow structures significantly [44,76]. dominant frequency increases.
The combined influences of the pressure ratio, impingement dis- The screech tone is generated through a feedback mechanism
tance and inclined angle on impinging flowfields of inclined jets have [88,89]. As illustrated in Fig. 9, instability waves are originated from
been evaluated by Fujii et al. [76] and Nakai et al. [44], using schlieren the lip of the nozzle exit and gradually grow downstream. Thence, the
graphs and pressure-sensitive paint/temperature-sensitive paint (PSP/ instability waves interact with the shock structures, producing the
TSP) graphs. The inclined impinging jet could be classified into three feedback waves that run upstream and stimulate the instability waves.
types according to the shape of plate shocks, where type II covers an The screech tone appears as a spike in the spectrum, with the dominant
variant type II′, as illustrated in Fig. 8. In type I, the plate shock re- frequency between the turbulent mixing noise and the broadband
sembles the Mach disk in the free jet and only one slipstream is ob- shock-associated noise. The first harmonic component of the screech
servable downstream of the lower tail shock. As the impingement dis- tone propagates in the direction of 90 to jet axis, while the dominant
tance decrease or the pressure ratio increase, h/ h s decreases and the components propagate in a direction larger than 90 . The screech tones
flowfield behaves as type II (h s is the axial distance from the nozzle exit usually occur along with strong oscillations, including toroidal mode

5
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

Fig. 7. Schlieren graphs of inclined impinging jets with different inclined angles. NPR = 3.5, h/ d = 1.5. Reprinted from Ref. [74], Copyright (2014), with permission
from Elsevier.

oscillations and helical/flapping mode oscillations. The dominant os-


cillation mode varies with the jet Mach number [90]. When the jet
Mach number is low (below 1.25 for the case in Ref. [90]), the domi-
nant oscillation mode is usually a toroidal mode, where the flowfield
and the acoustic field is symmetrical. As the jet Mach number increases,
the dominant oscillation mode could switch to a helical/flapping mode,
where the jet oscillates across a flapping plane and the flapping plane
itself is in a processional motion around the jet axis. It should be noted
that, the SPL peak of the screech tone goes down with increasing jet
temperature. As a result, the screech tone might be not observable in
some field measurements [23] or some scale model experiments
[1,3,21,22] of rocket launching.

2.2.2. Noise of normal impinging jets Fig. 9. Scheme of screech tone feedback loop.
As the flow oscillations accompanying with the screech tone in the
free jet [61], several oscillation modes could be observed in moderately
oscillation modes are observable: axisymmetric modes and helical
and highly under-expanded impinging jets [24,91,92]. Since the flow
modes. For an impinging jet with an axisymmetric mode, the flow is
structures of an impinging jet in the free jet region resemble those of a
usually symmetrical and the oscillation is in an axial or radial direction.
free jet [60], unstable flow in an impinging jet could show similarity to
For an impinging jet with a helical mode, the flow is usually asym-
a free jet under certain conditions. Apart from the noise generated by
metric and the oscillation is rotational. Under varying pressure ratios,
the oscillations in each region, i.e. the free jet region, impingement
impingement distances and inclined angles, variant oscillation modes
region and wall jet region, additional noise could be induced by in-
could be observed in the schlieren graphs [93], including helical modes,
teraction among the regions.
semi-helical modes, axially pulsing modes, axially flapping modes and
Experimental studies [93–95] showed that two main kinds of

Fig. 8. Classification of inclined impinging jets. θ is the angle between the jet axis and the plate surface, the complement angle of inclined angle. h is the impingement
distance. h s is the axial distance between the nozzle exit and the location where the first shock cell of the corresponding free jet under the same nozzle operational
conditions attains the maximum diameter. x is the axial distance between the nozzle exit and the point where the upper jet boundary intersects with the plate.
Reproduced with permission of ASME, from Ref. [76]; permission conveyed through Copyright Clearance Center, Inc.

6
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

axially collapsing modes, some of which are corresponding to those in


the free jet. In a helical mode, the location of the plate shock is almost
fixed, while the shock plane is inclined with a certain angle to the plate
and rotates around the jet axis. The semi-helical mode is a transitional
mode from axisymmetric modes to helical modes. In an axially pulsing
mode, the axial location and the diameter of the Mach disk oscillate
periodically. In an axially flapping mode, the plate shock axially flaps
upstream and downstream, just as a jellyfish. Besides, the normal shock
immediately upstream of the plate shock moves downstream and dis-
appears, before a new normal shock appears at the previous position
and continues the movement. The continual appearance and dis-
appearance of the normal shock are also observable in a free jet, where
the phenomenon is called the shock splitting [96]. In an axially col-
lapsing mode, apart from the normal shock moving downstream, a
normal shock moving upstream appears near the plate and collides with
the normal shock moving downstream at the end.
The oscillation of the flow structures observed in the impinging jet Fig. 11. Power spectrum density (PSD) of a free jet and impinging jets at
varying impingement distances. Reproduced with permission of AMERICAN
would simultaneously induce the oscillation of the pressure distribution
INSTITUTE OF AERONAUTICS AND ASTRONAUTICS, INC, from Ref. [100];
on the plate. Through the measurements with phase-conditioned
permission conveyed through Copyright Clearance Center, Inc.
schlieren and PSP, it is observed that the large-scale turbulent struc-
tures in the shear layer contribute to the oscillation of wall pressure
[94,95], as shown in Fig. 10. For an impinging jet dominated by an numerically. Still, it is difficult to predict the nozzle operational con-
axisymmetric mode, the wall pressure contour appears as concentric ditions where a specific mode would appear.
circles. while the pressure distribution resembles a yin-yang symbol The flow of the impinging jet in the free jet region show similarity to
when the flow is dominated by a helical mode. Both results are con- the free jet, for both stable flow [63] and unstable flow [24]. Similar
sistent with the linear stability analyses. flow induces similar noise, which is observable in a noise spectrum. As
As to numerical investigations on the stability of impinging jets, a shown in Fig. 11, when the impingement distance is large enough
global stability analysis [97] of a Large Eddy Simulation (LES) solution (h/ d = 10 in this case), the noise spectrum of the impinging jet closely
to the linearized Navier-Stokes (N–S) equations, showed that both ax- resemble that of the free jet. When the impingement distance decreases,
isymmetric (m = 0 ) modes and azimuthal modes (m = 1) are stable and the power spectrum density (PSD) peak of the screech tone decreases
only convective instabilities are present for m < 2, where m is the mode and finally the screech tone disappears. At the same time, another PSD
order. A Dynamic Mode Decomposition (DMD [98]) analysis [99] of an spike (the impinging tone) appears and the PSD peak value increases, so
LES solution to the linearized N–S equations revealed that there could does the dominant frequency of the impinging tone. It is confirmed by
be a threshold value for the disturbance wavelength. Disturbances of the experiment [101] that when the jet temperature increases, the
dimensionless wavelengths less than 1/8 are dampened immediately broadband noise is enhanced and the impinging tone is weakened for
while other disturbances grow in the streamwise direction. Also, in the the impinging jet, which is similar to the influence of the jet tempera-
shear layer exists a convective instability, that the wave packet splits as ture on the broadband noise and the screech tone for the free jet.
it moves downstream. The oscillation modes of impinging jets could be In addition to the noise similar to that in the free jet, there are
clearly observed and well analysed both experimentally and impinging tones induced by the feedback loop between the free jet

Fig. 10. Results of an axisymmetric mode


and a helical mode. The first row is for the
axisymmetric mode. The second row is for
the helical mode. The first column is for
phase-conditioned schlieren graphs. The
second column is for PSP graphs of the plate.
The third column is for contours of pressure
on the plate deducted from linear stability
analyses. Reproduced by permission from
Springer Nature Customer Service Center
GmbH: Springer Nature Experiments in
Fluids [94], Copyright (2015).

7
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

Fig. 12. Schlieren graphs of four phases in a feedback loop. Reproduced by permission from Springer Nature Customer Service Center GmbH: Springer Nature
Journal of Visualization [102], Copyright (2012).

region and the impingement region in the impinging jet. According to observable. As indicated in the review by Henderson [24], the im-
Powell [27], small disturbances originated from the lip of the nozzle pinging tones are generated by the oscillations near the wall. When a
exit grow as they progress downstream in the shear layer, producing Mach disk forms in the shock cell immediately upstream of the standoff
large-scale vortex structures convecting downstream. The impingement shock, the impinging tone is observable. However, when the nearest
of the large-scale structures on the plate produces pressure disturbance shock cell resembles an X shape, a stable standoff shock forms near the
waves running upstream, which stimulate the disturbances and close wall and the impinging tone disappears, which is defined as a zone of
the feedback loop, as shown in Fig. 12. silence in the impingement distance-NPR graph, as shown in Fig. 13.
Along with the theory presented by Powell [27], a prediction model Sinibaldi et al. [65] indicated that when the pressure ratio is below the
(equation (1)) of the impinging tone frequency is developed, where N is
the number of the vortices generated in one feedback period, hc is the
distance that the vortex convection covers, uc is the convection velocity
of the vortices, hs is the distance that the acoustic wave propagation
covers, Us is the average propagation velocity of the acoustic waves and
p is the phase lag between the acoustic feedback and the vortex gen-
eration.
N+p hc dh h
= + s
f 0 uc Us (1)

Weightman et al. [103] have discovered a periodic transient


shocklet structure in the wall jet region, which is a possible source of
the impinging tone. The life time of the structure explains the phase lag
term in Powell's model (equation (1)). Based on the findings, the pre-
diction model is revised to equation (2) and verified by the experiment.
N hc dh h Fig. 13. Zones of silence. is the first cell length, the distance from the nozzle
= + s + t exit to the end of the first shock cell in the corresponding free jet under the
f 0 uc Us (2)
same nozzle exit conditions. Reprinted with permission from Ref. [24]. Copy-
However, under certain conditions, impinging tones might not be right 2002, Acoustic Society of America.

8
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

zone of silence, the shear layer interacting with the plate would induce
the feedback towards the nozzle exit and thus generate the impinging
tone. When the pressure ratio is above the zone of silence, the im-
pinging tone is generated by the feedback between the standoff shock
and the plate. However, in the zone of silence, the standoff shock dis-
turbs the interaction between the shear layer and the plate, while the
standoff shock itself is not strong enough to produce a noticeable shock-
plate interaction. Henderson et al. [104] have observed oscillations of
the standoff shock, the recirculation region and the wall jet region.
Nevertheless, no explicit connection between the recirculation region
and the impinging tone is discovered. The impinging tone could be
detected for an impinging jet either with or without a stable re-
circulation region.

2.2.3. Noise of inclined impinging jets


When the inclined angle is small, only nuances of the flow structures
and acoustic characteristics between normal and inclined impinging
jets might be observed. As the inclined angle increses, the character-
istics of the noise spectra could show remarkable difference. Worden
et al. [105] have measured the nearfield and farfield spectra at 90 to
the jet axis, for normal and 40 -inclined impinging jets under varying
impingement distances and jet temperatures. For both kinds of im-
pinging jets, when the impingement distance increases, the nearfield
OASPL increases until it reaches the peak value at a certain impinge-
ment distance, while the farfield OASPL decreases monotonically with
increasing impingement distance. It might be attributed to the reflec-
tion of acoustic waves, which affects the nearfield measurement. The
SPL of the full band (1 kHz–30 kHz) measured either at nearfield or
farfield increases along with the increasing jet temperature. For the
normal impinging jet, under a fixed impingement distance, the im-
pinging tone appears in a certain range of jet temperature but dis-
appears when the jet temperature increases further, as shown in
Fig. 14a. For the inclined impinging jet, on the contrary, no impinging
tone is observed under the same conditions, unless the jet temperature
is high enough, as shown in Fig. 14b.
For the inclined impinging jet, the mechanisms of the noise gen-
eration and propagation have been widely studied either experimen-
tally [107–110] or numerically [46,106,111–117]. As a result, several
noise sources and influential factors have been revealed. Among the
numerical investigations, Nonomura et al. [111,112] studied the
acoustic field of the inclined impinging jet. Three types (I, II, III) of
acoustic waves could be observed under varying impingement distances
and jet temperatures, which is consistent with the computation by Fig. 14. Nearfield noise spectra of impinging jets at varying jet temperatures.
Brehm et al. [106], as shown in Fig. 15a. Type I is Mach waves induced Mj = 1.5, NPR = 3.7 , h/ d = 6. Reproduced by permission from Theodore J.
by the shear layer of the main jet. Type II is acoustic waves originated Worden [105].
from the impingement region. Type III is acoustic waves induced by the
shear layer in the wall jet region. For type II, when the impingement coefficients (see Fig. 15b), as well as acoustic waves radiating in an-
distance increases, both the SPL and the dominant frequency decrease, other direction. For the octave band centered at St = 0.16, only the SPL
while the propagation direction remains almost unchanged. When the of type III acoustic waves is relatively large, which is consistent with the
jet temperature increases, the shear layer of the main jet becomes PSD peak near St = 0.2 observed 75 to the jet axis. For St = 0.33, the
thicker, the length of potential core decreases, the SPL of the full-band, type II acoustic waves are also noticeable in the SPL map. For even
especially the low frequency components, increases and the propaga- larger Strouhal numbers (St = 0.66, 1.31), the SPLs of type II and type
tion angle to the jet axis decreases. III are reduced, which is consistent with the PSD peak near St = 0.5
To further analysis the three types of acoustic waves, Nonomura and observed 120 to the jet axis. Through the causality method, which has
Fujii [113,114] and Tsutsumi et al. [115] have adopted the short time been applied to the identification of noise sources in turbulent free jets
Fourier transform (STFT) to extract sampling sequences at character- [119–121], Brehm.
istic frequencies for the numerical results of an inclined impinging jet, et al. [106] also confirmed that type II acoustic waves are originated
which are then analysed by the Proper Orthogonal Decomposition (POD from the impingement region and type III the wall jet region.
[118]). Two kinds of acoustic waves could be observed in the initial Two sources contributing to type II acoustic waves are discovered
three orders of POD modes. The main noise source (type III) is down- [46,116]. The first one is the interaction between the main jet shear
stream of the impingement region and the corresponding acoustic layer and the plate shock. The second one is the interaction between the
waves propagate in a small angle to the plate surface. The secondary main jet shear layer and the separation shock induced by the stagnation
noise source (type II) is located where the shear layer and the plate bubble. These two sources are easier to distinguish at a relatively small
intersect, and radiates the acoustic waves nearly along the plate inclined angle, while the acoustic waves generated by the two sources
normal. In the POD analyses conducted by Brehm et al. [106], the three are not distinguishable at farfield. It is found that at a smaller inclined
types of acoustic waves are observable in the phase plot of Fourier angle, the plate shock is stronger [117], the OASPL is higher and the

9
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

Fig. 15. (a) Isocontour of Q-criterion, Qd/ uref = 1.2 × 105 and contour of normalized gauge-pressure. (b) Phase of complex Fourier coefficients, St = 0.66 . Reproduced
with permission of Cambridge University Press, from Ref. [106].

dominant frequency is lower, however, the mechanism how the in- platform, which becomes the dominant noise source again. As the
clined angle influences the noise spectrum is still unclear. rocket ascends further, the exhaust tends to be a free jet and the exhaust
The acoustic waves identified in the numerical investigations above itself is the dominant noise source. The results of scale model experi-
are also confirmed by experiments. Akamine et al. [107,108] observed ments [1,3,41] of rockets at different lift-off heights led to consistent
the acoustic waves propagating in two different directions, corre- conclusions.
sponding to type II and type III, as shown in Fig. 16a. However, no In addition, it is indicated in the 1/40 scale model experiment of
connection between the intensity of the plate shock and the OASPL of Ariane 6 [41] that different noise sources are dominant in different
type II is observed. Akamine et al. [109] attained three types of acoustic frequency bands during rocket launching. For a moderate lift-off height
waves using phase average schlieren, where the types of acoustic waves (15d ), noticeable noise is observed at the engine exhaust, the launch
detected are different at different impingement distances. Further study platform, the exit of the deflector duct as well as the intersection be-
by Akamine et al. [110,122] shows that one of the three types of tween the deflector duct and the ground (the duct is partially under-
acoustic waves observed is reflected waves of type I, as shown in ground). In the octave band centered at 0.5 kHz (St = 0.01875), the
Fig. 16b. The variation of the impingement distance leads to the change engine exhaust, the launch platform and the duct exit are dominant
of area covered by the reflected waves, and thus the change of noise noise sources. As the frequency increases, the SPL of the noise observed
spectra measured at certain fixed points. from the duct exit and the engine exhaust goes down and almost dis-
In summary, three types of acoustic waves could be observed in the appears in the octave band centered between 4 kHz and 16 kHz
inclined impinging jet, as illustrated in Fig. 17. The first type is Mach (0.15 < St < 0.6 ). Simultaneously, the duct intersection becomes a sec-
waves originated from the shear layer of the main jet (I) and the cor- ondary noise source, while the launch platform remains the dominant
responding reflected waves (I′). The second type is acoustic waves in- noise source. By analyzing the deconvolued density spectral level of
duced by the interaction between the shear layer and shocks, including different noise sources, it is found that the dominant noise source is the
the interaction with the plate shock (II) and the interaction with the duct exit at h = 0 . As the lift-off height increases to 10d , the duct exit is
separation shock (II’). The third type is Mach waves generated by the the dominant source of noise with frequancy below 4 kHz (St < 0.15)
shear layer of the wall jet (III). while the launch platform is the dominant source of noise with fre-
quancy above 4 kHz (St > 0.15). At h = 20d , the launch platform is the
dominant source of noise in full-band observed (250Hz < f < 16kHz
3. Impinging jet interaction with launch structures 0.009375 < St < 0.6 ) and the duct intersection becomes a secondary
noise source for f > 2kHz (St > 0.075). When the lift-off height reaches
Rocket launching is a dynamic process, including engine ignition 30d , the launch platform is the dominant source of noise with frequancy
and rocket lift-off. The dynamic effects during the engine ignition are below 1 kHz (St < 0.0375) while the duct intersection is the dominant
mainly induced by the development of the engine exhaust from the source of noise with frequancy above 1 kHz (St > 0.0375).
ignition to the steady operation of the engine. While during the rocket
lift-off, the lift-off velocity is much lower than the jet velocity, the
dynamic effects are mainly induced by the change of lift-off height. 3.1. Interaction with the launch platform
During different phases of rocket launching, the impinging effects of
the engine exhaust are dominated by its interaction with either the The launch platform is generally set beneath the rocket, with several
launch platform or the deflector system [21–23]. Immediately after the cut-outs on the launch platform, where the engine exhaust passes. The
engine ignition, the engine exhaust develops and impinges on the noise induced by the impingement of the engine exhaust on the de-
launch platform dynamically. The launch platform is the dominant flector would be partially sheltered by the launch platform. However,
noise source. After the engine operates steadily, the exhaust passes the noise originated above the launch platform would be partially re-
through the launch platform cut-out and gets released from the exit of flected and become additional acoustic loads on the nose fairing. When
the deflector duct. The duct exit becomes the dominant noise source. the launch platform is directly impinged by the engine exhaust, the
During the lift-off, the exhaust grows and impinges on the launch launch platform itself would also become a noise source [21–23,41].

10
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

platform and a critical lift-off height where the exhaust starts to im-
pinge on the launch platform. As the lift-off height reaches the critical
value, the impingement on the launch platform becomes a noise source.
The acoustic characteristics of the exhaust impinging on the launch
platform are similar to that of the jet normally impinging on a plate,
which is almost irrelevant to the size of the cut-out [42].
In the numerical investigation [123], Reynold number (Re) is ad-
justed to control the state of the normal impinging jet on plates with or
without a cut-out. When the jet completely passes through the cut-out,
the nearfield OASPL of the noise radiated from the plate with a cut-out
is 30 dB lower than that without a cut-out. When the jet partially im-
pinges on the plate, the maximum overall sound reduction level (SRL) is
about 10 dB, but the overall SRL values probed at most points are lower
than 5 dB, which indicates that the noise reduction effects of the cut-out
would be limited once the plate gets impinged.
The effects of the cut-out size have been studied through scale
model experiments [42]. Under a lift-off height smaller than about 10d ,
the mixing layer of the rocket exhaust passes through the launch plat-
form exactly with a cut-out of 1.58d . The OASPL of the acoustic loads on
the nose fairing received from the launch platform is lower than either
the launch platform without a cut-out or with a larger cut-out (2.46d ).
The above results indicate that the cut-out helps to avoid the noise
originated from the jet impingement on the launch platform, but a large
cut-out would weaken the shelter effect on the noise induced by the jet
impingement on the deflector. As the lift-off height grows to 10d , the
dominant frequency of the peak of 1/3 octave spectral level is shifted
from St = 0.5–0.9 to St = 0.1–0.4. The radiation peak line for the oc-
tave centered at St = 0.52 (8 kHz) is 125 to the jet axis and that for the
octave centered at St = 0.13 (2 kHz) is 170 to the jet axis, which shows
that the low frequancy components tend to propagate along the rocket.
When the lift-off height reaches 13.9d , the effects of the cut-out on the
OASPL measured at the nose fairing become negligible, which is con-
sistent with the results of the scale experiment by Malbqui et al. [41].
Thus, normal impinging jets on a plate could serve as reasonable re-
ferences for the exhaust impingement on the launch platform during
this phase. As the lift-off height increases further, both the energy ra-
diated from the mixing layer and reflected by the launch platform in-
creases, and thus the OASPL continues to increase. After the lift-off
height surpasses the length of jet potential core, the mixing layer is
almost completely developed and the OASPL starts to decrease for that
Fig. 16. Schlieren graph and conditional sampling of an inclined impinging jet. the decrease of energy reflected by the launch platform exceeds the
Figure (a) is reproduced by permission from Masahito Akamine [108]. Copy- increase of energy radiated from mixing layer.
right 2015. Figure (b) is reproduced by permission from Masahito Akamine, The acoustic fields under different lift-off heights could be observed
Koji Okamoto, Kent L. Gee, Tracianne B. Neilsen, Susumu Teramoto, Takeo in the computation by Tsutsumi et al. [43], which is shown in Fig. 18.
Okunuki, and Seiji Tsutsumi [122]. Copyright 2018. The noise is probed near the nozzle exit with 90 to the jet axis. At a lift-
off height of 6d , the exhaust is relatively stable and almost passes
through the cut-outs completely, leading to a low SPL peaked at
St = 0.1. When the lift-off height reaches 11d , the exhaust becomes
unstable and impinges on the launch platform, resulting in a higher
OASPL, while the peak SPL remains near St = 0.1. As the lift-off height
increases to 16d , the mixing layer develops, the Mach waves radiating
downstream and the corresponding reflected waves are observed. The
probed OASPL continues to increase and the peak SPL shifted to
St = 0.04. When the lift-off height is finally 21d , more Mach waves
radiating downstream are observed, whereas the reflected waves are
greatly weakened. The OASPL decreases and the peak SPL is near
St = 0.03.
Based on the understanding of the noise generation of the launch
Fig. 17. Three types of acoustic waves. platform, two kinds of passive methods to reduce the noise are under
research. One is to apply sound absorption treatment to the surface of
the launch platform. The SPL of the full band could be reduced by
During different phases of rocket launching, the dominant effect of
setting a porous plate between the nozzle exit and the impinged plate
the launch platform varies. When the lift-off height is small, the exhaust
[124,125]. Also, the surface roughness of the impinged plate is posi-
completely passes through the launch platform cut-out. The noise is
tively correlated to the reduction of the turbulent noise [126]. The
mainly induced by the impingement of the exhaust on the deflector.
other kind of method is to perforate the launch platform in addition to
The size of the cut-out influences the shelter effect of the launch
the main cut-outs. In the scale model experiments [3], for the

11
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

Fig. 18. Acoustic field of H-IIA at different lift-off


height. The black line at the bottom indicates where
the launch platform is located. The launch platform
cut-outs and the deflectors below are not visible in
the figure. Reproduced with permission of AMERI-
CAN INSTITUTE OF AERONAUTICS AND ASTRON-
AUTICS, INC, from Ref. [43]; permission conveyed
through Copyright Clearance Center, Inc.

configuration without a launch platform, the OASPL measured on the 3.2.1. Single-sided deflector
rocket and 13d above the nozzle exit is at least 4 dB lower than that with Single-sided deflectors direct the engine exhaust to one side. These
a launch platform. For the perforated launch platform, the OASPL deflectors have been applied to rocket launching, mobile missile
measured is in between and the most sound reduction is observed in launchers [129] and conventional takeoff and landing (CTOL) carrier-
low and medium frequency bands. As the lift-off increases, the noise based aircrafts [130–132]. The simplest deflector is an inclined plate,
reduction becomes more significant. The surface treatment mainly in- which has been studied in many fundamental researches. For CTOL
creases the dissipation of the acoustic disturbances while the perfora- carrier-based aircrafts, inclined plates are usually utilized as jet blast
tion is to increase the acoustic transmittance. Based on the study by deflectors. For mobile missile launchers, planar or curved plates are
Gaeta et al. [127] and an unpublished report by Georgia Tech on sound utilized. However, for rocket launching, apart from a carefully designed
absorption material, Ahuja et al. [128] combined the two methods curved deflector surface, the deflector is usually extended by a deflector
above and proposed a design where the launch platform and the ground trench. For a closed deflector, the trench would be partially covered by
are covered by the sound absorption material while the cut-outs on the a plate, or be replaced by a duct.
launch platform are replaced by the perforation. The report by Georgia Design of an open deflector focuses on the optimization of the de-
Tech indicated that the sound absorption material greatly reduces the flector shape, including the surface near the impingement region and
medium and high frequency components of the noise. The study by the transition from the deflector to the trench and other structures.
Gaeta et al. [127] shows that the perforation could increase the noise in Tsutsumi et al. [133] numerically investigated the M − V rocket ex-
the high frequency band while reduce greatly the noise in thermal low haust impinging on a bucket-shaped deflector. Acoustic waves propa-
and medium frequency bands and eliminate the screech tone. gating in two directions are observed, which indicates two dominant
noise sources. One is located in the impingement region, where the
unsteady motion of the tail shock and the separation bubble generates
3.2. Interaction with the deflector the noise, propagating towards the rocket and inducing most of the
acoustic loads. The other is located at the edge of the deflector, where
The deflector system is designed to deflect the engine exhaust, re- the flow leaves the deflector in a flapping motion. The corresponding
ducing and eliminating backflow, and thus reduce the acoustic loads on acoustic waves are relatively stronger while the propagation is away
the rocket. There are several kinds of deflectors, including closed de- from the rocket. When the deflector shape is revised to a flat plane,
flectors with a closed duct/trench and open deflectors without one. though the separation bubble is still there, the acoustic waves propa-
According to the main deflection direction, deflectors could be divided gate away from the rocket, and thus the acoustic loads are reduced.
into single-sided and multiple-sided deflectors. Different deflectors Tatsukawa et al. [134–136] utilized numerical simulation to opti-
show different acoustic efficiency (sound power/exhaust's mechanical mize deflectors. As shown in Fig. 19, by keeping the impingement
power) [31]. A bucket-shaped deflector (with a deflector duct) shows distance constant, three type of deflectors are designed after several
high acoustic efficiency above 0.15, while a cone-shaped one shows a iterations: 1) the basic deflector, a 45 inclined plate as the baseline, 2)
low efficiency about 0.05 [31]. A 45 inclined plate shows an inter- the balanced deflector, where there is a bump upstream of the im-
mediate efficiency of 0.10. For large rockets, closed deflectors are pingement region to obtain considerable noise reduction with minimum
usually adopted, while for smaller rockets, a planar or curved surface is change of the shape, 3) the optimum deflector, where there is a bump
often the choice [16]. upstream of the impingement region and a sink downstream of the

12
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

Fig. 19. Acoustic fields of impinging jets on deflectors with different shapes. Reproduced by permission from Tomoaki Tatsukawa [134], Copyright 2016.

impingement region to minimize the acoustic loads. All the three types weaken the ignition overpressure. The unsteady simulation of the H-IIB
of acoustic waves introduced in Sec. 2.2 could be observed for the basic launch by Tsutsumi et al. [142] revealed the development of the
deflector. However, the balanced deflector weakens the type II acoustic flowfield after engine ignition. In the beginning, the leading shock is
waves by about 10 dB and the optimum deflector almost eliminates the generated by the ignition overpressure. Then the leading shock pro-
type II acoustic waves. Considering the flow structures, there are plate pagates outwards and gets reflected by the deflector. The shelter effect
shocks and separation bubbles for all the three deflectors, while the size of the launch platform and the deflector cover plate weakens the
of the separation bubble is the smallest for the optimum deflector, and leading shock and reduces the aerodynamic loads on the rocket. Still,
the largest for the basic deflector. Through the optimization, the local there are noticeable pressure disturbances. The starting vortices form
inclined angle in the impingement region is increased, resulting in a behind the leading shock and attach downstream of the impingement
smaller separation bubble, which weakens the separation shock and the region. Then the vortices gradually grow in size and fill the deflector
interaction with the shear layer, and thus reduces the noise. On the trench, finally induce secondary vortices at the trench exit. The simu-
contrary, when the inclined angle is small, strong recirculation and lation of H-IIA launch by Tsutsumi et al. [43] shows that when the lift-
separation shocks could appear [137]. However, by comparison be- off height is 16d, almost all the flow in the deflector trench is subsonic.
tween two deflectors resembling the basic one and the balanced one The acoustic waves propagate along the trench, and get reflected and
respectively [46], the spectra observed about 120 to the jet axis and 20d diffracted near the trench exit. The acoustic field outside the trench is
to the impingement point are almost the same for the two deflectors, similar to the superposition of spherical waves.
and no reduction is observed for the SPL peak (St 0.4 ) and the sec- However, the structures of closed deflectors suffer more from the
ondary SPL peak (St 5) in the spectra observed near the nose fairing. thermal loads induced by the engine exhaust. The different configura-
Still, the balanced deflector reduces the SPL by about 15 dB for tions of the deflector cover plate and the deflector trench are analysed
1 < St < 4 and St > 5. For practical optimization of deflector system, the through the simulation of a KSLV-II launch [143]. The results show that
object to reduce the SPL in a desired frequancy band might lead to a the deflector cover plate hinders the heat exchange between the engine
better optimization than to reduce the OASPL received at nose fairing. exhaust and the surrounding fluid. Thus, the average temperature in
The optimization of deflectors is also bounded by the terrain in deflector trench is relatively higher. Compared with a straight deflector
practical applications. Hence, the transition from the deflector to the trench, a trench converging near the trench exit induces additional
trench or the terrain should be carefully designed. Tsutsumi et al. [138] compression and increases exhaust temperature near the trench exit.
numerically investigated the M − V rocket exhaust impinging on an Besides the trench/duct downstream of the deflector, there is a
open deflector. Under fixed inclined angle and impingement distance, configuration where a duct is set upstream of the deflector. Tsutsumi
an ideal plate deflects the acoustic waves away from the rocket, while et al. [145] have studied this configuration by computation. The con-
the deflector with a step transition induces additional acoustic waves figuration is similar to the Launch Pad 0B at NASA's Wallops Flight
originated near the transition, and the deflector with an arc transition Facility, as shown in Fig. 20. A duct upstream of the deflector connects
almost eliminates the additional acoustic waves. a small launch platform and a deflector. Arc transition is adopted and
Tsutsumi et al. [139] numerically compared the flowfields and no duct is extended downstream of the deflector. As the impingement
acoustic fields of the rocket exhaust impinging on two different de- distance increases, the OASPL increases at first and decreases after
flectors. One is a 45 inclined plate, the other is a 45 inclined plate with reaching the peak value. The trend is similar to the conclusions in Sec.
an arc transition, which is designed to eliminate the tail shocks. All the 3.1.
three types of acoustic waves are observable for both deflectors, but the Three different deflectors with a duct upstream are compared
tail-shock-free deflector avoids additional shock structures induced by through scale model experiments [146]. The first one is a basic de-
the deflection of the wall jet and thus the corresponding noise. flector similar to that in Fig. 20 by replacing the trench with a duct. The
Though open deflectors are simple and easy to implement, large second one is the basic deflector with a duct extended downstream of
rockets usually require closed deflectors to obtain sufficient noise re- the exit. The third one is based on the second deflector and the duct exit
duction [16]. In a comparison of the sound reduction effects between shape is revised to a flat rectangular. At a lift-off height of 8d , both the
open deflectors and closed deflectors [140,141], the OASPL is measured measurements near the nozzle exit and the nose fairing led to a con-
at several points evenly distributed on a horizontal circle of radius 80d . sistent conclusion that the full band SPL for the second deflector with a
Compared with the free jet noise in an azimuth range of 0–90 (0 for the extended duct is lower than that for the basic deflector, while the
designed deflection direction), for an open deflector, the OASPL mea- OASPL for the third deflector with a flat duct exit is higher than that for
sured in the range of 70 –90 is lower, while the range is extended to the basic deflector. The peak OASPL is about 160 to the jet axis for the
45 –90 for a closed deflector, with the SRLs larger than the open de- first two deflectors, while for the deflector with a flat duct exit the
flector. Thus, the directivity of the closed deflector is better than the dominant direction of propagation is reduced to 150 . The shape of the
open deflector. Also, a scale model experiment of the Ariane 5 [51] duct exit has noticeable influence on the directivity and the SPL re-
rocket indicated that the extend of closed deflectors could reduce the ceived by the nose fairing.
noise effectively. The Uchinoura space center adopted a deflector resembling the
In addition to noise reduction effects, closed deflectors help to second deflector, the basic deflector with a duct extended downstream

13
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

service structure, the wavelet power spectra also show a dominant


frequancy of about 250 Hz when h < 12d . As the lift-off height increases
further, the service structure is affected directly by the engine exhaust
and the wavelet power spectra show a broadband noise (30 Hz–4 kHz,
0.02 < St < 2.67 ).
To summarize, the design of the deflector surface, the deflector
duct/trench, as well as the transition in between, is strongly associated
with the local flow structures and the sound generation and propaga-
tion. The single-sided deflector is a simple configuration, but the con-
figuration covers all the design issues above, which are also concerned
with the studies on the multiple-sided deflector.

3.2.2. Multiple-sided deflector


Multiple-sided deflectors scatter the engine exhaust to multiple di-
rections. In practical applications, wedge-shaped deflectors are more
usual than cone-shaped deflectors. The wedge-shaped deflector could
function as a single-sided deflector when only one side of the wedge is
impinged by the engine exhaust. In this situation, the flow structures
may resemble the inclined impinging jet on a plate or a single-sided
deflector. However, if the jet impinges near the tip, complex flowfields
could appear at the tip and show different characteristics, which is only
investigated by a few fundamental studies [148–154] and numerical
simulations [155,156].
The experiments of impinging jets on wedges and cones [148,149]
have shown that, when the wedge or cone tip is near the location of the
Mach disk in the free jet under the same nozzle operational conditions,
the Mach disk in the impinging jet would be pushed upstream, which is
similar to the trend in the normal impinging jet. Lamont and Hunt
[150] have studied impinging jet on wedges under varying jet Mach
numbers, pressure ratios, impingement distances and wedge angles. As
Fig. 20. MARS launch pad 0B. Reproduced from Ref. [144]. Courtesy NASA/
JPL-Caltech.
the pressure ratio increases, the shock structures near the wedge tip
would successively transit from a detached shock to an attached shock,
then double triple-point structures and finally four-shock confluence
of the exit. The flight data of the Epsilon-1 launch are collected by the structures, as shown in Fig. 21. The critical transition pressure is pre-
microphones inside the nose fairing and on the service structure near dicable through shock polars. However, a detached shock could appear
the nose fairing [147]. The OASPL measured inside the nose fairing while the cone shock theory predicts an attached shock [151].
reaches the peak value 132.2 dB for a lift-off height between 6d –7d . The As to the aerothermodynamic and acoustic effects of impinging jets
wavelet power spectra of the noise inside the nose fairing show that the on wedges or cones, the experiments by Hunt [152] have shown that
dominant frequancy is about 250 Hz (St 0.167 ) before the lift-off the aerodynamic load on the deflector is nearly independent of the
height reaches 12d , while the dominant frequancy is shifted below impingement distance and the pressure ratio, but relevant to the wedge
125 Hz (St 0.083) when h > 12d . As to the acoustic loads on the angle. Vinze et al. [153] have studied the thermal effects of impinging

Fig. 21. Shadowgraphs of impinging jets on a wedge at different pressure ratios. Designed Mj = 2.2 , h/ d = 1. (a) Detached shock. PR = 0.8. (b) Attached shock.
PR = 1.0 . (c) Double triple-point structures. PR = 1.4 . (d) Four-shock confluence structures. PR = 1.5. Reproduced with permission of Cambridge University Press,
from Ref. [150].

14
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

products of the solid propellant combustion include a large number of


discrete solid (or molten) particles, which are extra heat sources for the
exhaust [28]. The liquid and solid particles could exert a significant
influence on the exhaust. Thus, it is necessary to consider the influence
of the particles when trying to investigate the engine exhaust im-
pingement during rocket launching.
Numerical methods for the multiphase flow and the discrete phase
flow have been summarized in many reviews [157–163]. Numerical
methods for the discrete phase flow could be categorized into Eulerian
methods or Lagrangian methods. The Eulerian methods treat discrete
particles as a continuous phase and then adopt numerical methods for
the multiphase flow, while the Lagrangian methods directly track the
particles. In general, the volume fraction of solid particles in the ex-
haust of a typical rocket engine is relatively low [28]. Thus, the La-
grangian methods are applicable, while the Eulerian methods are
usually considered less suitable for the simulation of particles in the
engine exhaust. Troyes et al. [29] adopted an Eulerian method and a
Lagrangian method respectively to simulate a solid rocket exhaust
containing solid particles of two sizes. Noticeable inconsistency is found
between the results of these two methods, especially for the solid par-
ticles of a relatively large size.
The water injected is usually in a continuous phase, while the water
jet could interact with the impinging jet break up, and become ato-
mized. Thus, modelling of the liquid atomization [164] is required to
simulate the impinging jet interaction with aqueous jets. In addition, it
is indicated that the atomization of the water jet is related to the tur-
bulent disturbances. The experiment by Ragaller et al. [54] has been
simulated through an unsteady RANS (URANS) method, and a hybrid
URANS-LES method respectively, in conjunction with a VOF (volume of
Fig. 22. Typical configurations of a water injection system. Reproduced by
fluid) method [47]. The results indicate that the URANS method could
permission from SAGR Publications [20].
hardly simulate the atomization of the water jet, whereas the hybrid
method is capable of capturing the phenomenon. However, high-fide-
jets on wedges under different pressure ratios, impingement distances lity turbulence simulation and particle tracking require costly compu-
and wedge angles. The results show that when the impingement dis- tations, which makes it necessary to balance the computational re-
tance is lower than 6d , the recovery factor at the stagnation point is sources for capturing the turbulent structures and tracking the particles.
almost independent of the wedge angle, and varies in a range of By choosing limited particles to simulate with proper strategies [165]
0.85–0.95 under varying nozzle pressure ratio between 2.36 and 5.08. and reducing the number of particles tracked, it is possible to save the
Also, the wedges show higher heat transfer rate than the flat plates. computational cost.
Patel and Mathew [154] have compared the flowfields and the noise of
two-dimensional jets impinging on wedges under varying impingement 4.1. Interaction with aqueous jets
distances and wedge angles. At an impingement distance about 1d , the
impinging jet is steady. As the impingement distance reaches about 4d , During the operation of the water injection system, the liquid phase
large oscillations are observed in the flowfield and the OASPL is about exists in forms of either the continuous aqueous jet or the discrete liquid
7 dB higher. Under the same impingement distance, the wedge with a particle. As to the momentum exchange with the liquid phase, the ex-
larger wedge angle induces a detached shock and the OASPL is 3 dB change between the aqueous jet and the engine exhaust could decele-
higher than that with an attached shock. rate the exhaust and change the velocity profile [20,51,166–170], while
For studies of engine exhaust impinging on the deflector system, the the exchange with the liquid particles could reduce the turbulent in-
rocket exhaust impinging on three configurations of deflector, a flat tensity of the exhaust [171–174]. In addition, the vaporization of the
plate, a wedge deflector or a cone deflector set in a trench has been liquid phase contributes to considerable heat absorption. Both the
simulated [156]. Both simulations of rockets with two and six nozzles momentum and heat exchange result in noise reduction.
show that the thermal loads on the cone deflector are lower than those As illustrated in Fig. 22, there are several configurations of the
on the flat plate and the wedge deflector. Besides, the cone deflector is water injection system. The water injection system could be set near the
more suitable for the rocket with six nozzles than that with two nozzles. nozzle exit, around the cut-out on the launch platform, near the de-
flector top, on the deflector surface or near the deflector trench exit.
4. Impinging jet interaction involving multiphase flow Water injections at different locations have varying effects. The scale
model experiments of Ares I [21,22,41] showed that the water injec-
Commonly, there are two kinds of interaction with impinging jets tions located near the launch platform cut-out, the deflector top and the
involving multiphase flows during rocket launching. In general, a water deflector trench could effectively reduce the noise induced by the de-
injection system is included in a launch pad for large vehicles to cool flector, while the water injections located on the launch platform top
down the launch system and reduce the noise. The water jet could be could reduce noise originated from the impingement region and limit
atomized by the engine exhaust [47], producing a large number of the extent of the noise.
discrete droplets. Through vaporization, the water jet and the droplets
absorb heat and cool down the exhaust and the launch structures. Si- 4.1.1. Aqueous jet for heat reduction
multaneously, the exhaust is decelerated by the momentum exchange The heat reduction effect of the water injection mainly contributes
with the liquid phase. The cooling and deceleration of the exhaust could to the heat protection of the impinged structures, including the launch
reduce the noise of impinging jet [1,20]. For a solid rocket engine, the platform and the deflector system. The effect is determined by the

15
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

configuration, and the characteristic parameters of the injection as well impinging jet of different jet temperatures. Ignatius et al. [1,20] have
as the engine exhaust. The conditions of the engine exhaust are usually found that the noise reduction effects for the hot jet are better than the
given by the desired launch vehicles. Before the study of the injection cold jet at different lift-off heights tested. When the rocket is near the
conditions, the configuration should be determined firstly. In a nu- launch platform, the injections reduces the noise of the full band, where
merical investigation of the water injection [175], the engine exhaust the main reduction is in the medium frequency band for the cold jet
impinges on one side of a wedge deflector, and water is injected on the while additional reduction is observed in the high frequency band for
top of the deflector in a horizontal direction towards the exhaust. The the hot jet. However, the results of the experiments by Ragaller et al.
results showed that most water is vaporized before reaching the de- [54] lead to a different conclusion, that the impinging noise in the high
flector surface and part of the water jet moves away from the exhaust, frequency band is enhanced by the water injections at higher jet tem-
resulting in a limited heat reduction effect on the deflector surface. A perature. When the impingement distance is 8 times of the nozzle throat
numerical and experimental investigation [30] showed that certain diameter, the impinging tone appears and the injection reduces the SPL
configurations of water injection system near the nozzle exit could in- peak of the impinging tone. For a temperature ratio (stagnation tem-
hibit the afterburning of the engine exhaust, thus effectively reducing perature at the nozzle inlet/ambient temperature) of 1.0 or 2.04, the
the thermal loads and the size of the ablation region. The injection noise in the high frequency band is reduced as well. On the contrary, for
conditions of a water injection system located on the deflector surface a temperature ratio of 2.81, noise of the high frequency band is en-
has been studied experimentally [55]. The maximum wall temperature hanced. When the impingement distance is 10 times of the nozzle throat
with no water injection is 3522 K while that with the water injection is diameter, the results are similar for a temperature ratio of 1.0 or 2.04,
2777 K. By taking the total injection flow rate fixed, as the number of while for a temperature ratio of 2.81 or 3.54, no impinging tone is
injection holes increases and the size of injection holes decreases, observed and the noise of the full band is enhanced, especially the noise
smaller water particles are produced and the maximum wall tempera- in the high frequency band.
ture drops to 1311 K. Under a constant injection flow rate, smaller in-
jection holes are preferred for the heat reduction. 4.2. Interaction with discrete particles

4.1.2. Aqueous jet for noise reduction The discrete solid (or molten) particles in the engine exhaust come
The noise reduction by the aqueous jet comes from the heat re- from the combustion of the solid rocket propellants. As indicated in the
duction and the deceleration of the exhaust, which reduce the energy of review [28], the alumina is molten under typical conditions in the
both the main jet and the turbulent structures. Just as the heat reduc- combustion chamber, but in the exhaust, the alumina temperature is
tion is affected by the aqueous jet, the configuration of the injection near the melting point and could be solidified downstream. The gas
system also has remarkable influence on the noise reduction effects. The phase and the discrete phase in the exhaust are usually not in equili-
scale model experiments of Ariane 5 rocket [51] showed that, under a brium. The nonequilibrium is more significant for the exhaust exiting
constant flow rate, the injection near the entrance of the deflector from a short nozzle, where the alumina temperature and phase are
system could suppress the noise originated from the deflector better more sensitive to the particle size and location. Rodionov et al. [48]
than other configurations. Ignatius et al. [1,20] have conducted a scale simulated a rocket exhaust, where the afterburning of the exhaust and
model experiment, where the exhaust of a rocket with two nozzles the phase change of the alumina particles are considered, including two
impinges on the either side of a close wedge deflector respectively. The variants of alumina crystalline structures, meta stable gamma and
results indicate that the injections near the nozzle exit are preferred to stable alpha. The results showed that the phase change of alumina
reduce the acoustic loads. As the lift-off height increases, the noise re- crystalline structures has negligible effect on the aerodynamic char-
duction by the injections near the launch platform and the deflector acteristics, but the phase change influences the optical properties and
decreases, while the injection set at the corresponding height could thus the radiation characteristics.
provide noticeable noise reduction. The distribution of the sizes of the alumina droplets in the chamber
The water pressure, flow rate and size of injection holes are coupled is bimodal, with small micron-sized smoke droplets originating from
parameters to characterize the injection conditions. Ignatius et al. combustion of the aluminum vapor and large droplets from the het-
[1,20] have studied these parameters by controlling the hole size and erogeneous combustion of the original aluminum globules of 20–200
the water pressure independently. With a fixed size of the injection µm [28]. When crossing the nozzle throat, due to the Rayleigh-Taylor
holes, the flow rate is determined by the water pressure and the highest instability, the large alumina droplets break up for the shear force on
SRL could be obtained at an optimum water pressure. The spectrum the order of 106g (g is the gravitational acceleration) [176,177]. Thus,
analyses showed that, when the water pressure is higher than the op- the particles collected from the exhaust are generally micron-sized and
timum value, the noise in the high frequency band is reduced more, but seldom larger than about 10–15 µm . Considering the break-up of the
additional noise is observed in the medium frequency band, which particles, prediction models of the particle diameter distribution based
might be induced by the impact of the water jet and become more on the critical Weber number are proposed. There are also distribution
noticeable as the water pressure increases. Under a fixed water pres- models based on statistics [178], including unimodal models such as
sure, the size of injection holes is changed to control the flow rate. The log-normal, Rosin-Rammler and power law-exponential laws, mean-
trend of the SRL varying with the flow rate resembles an S curve, where while, there are experiments showing multimodal distributions
the SRL surges in a certain range of mass flow rate while the SRL grows [179,180].
slowly outside the range. The trend is not the same for all the injection The particle size and the particle loading (mass flux ratio, mass flux
system configurations. For the injections at the launch platform bottom, of the particles/mass flux of the gas) are two primary parameters re-
the trend of the SRL behaves as an S curve when the lift-off height is 0, levant to the flow involving discrete particles. Dash et al. [181] re-
while the SRL quickly reaches the limited value of 2 dB at a lift-off viewed the studies on the two-phase flow concerning rocket exhaust
height of 8d. However, for the injections near deflector duct cover and summarized three critical parameters, the nondimensional relaxa-
plate, the SRL remains about 1 dB at all the lift-off heights tested. The tion length, the nondimensional relaxation time and the mass fraction
influence of the impingement distance on the injection configurations is ratio, as shown in Table 1. The first two parameters are related to the
also observable in the scale model experiment by Malbqui et al. [41]. particle size. For smaller particles, the relaxation length is smaller and
The results show that the injection at the duct entrance is effective the relaxation time is shorter, and thus the exchange of momentum and
before the lift-off height reaches 5d , while the SRL of the injection on energy between the two phases is more sufficient and it is easier to
the launch platform is noticeable for 5d < h < 30d . attain the flow equilibrium. However, at the nozzle exit, there could be
The noise reduction by the water injection is different for the considerable nonequilibrium between the particles and the gas phase

16
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

Two phase interaction parameters. Reproduced with permission of AMERICAN INSTITUTE OF AERONAUTICS AND ASTRONAUTICS, INC, from Ref. [181]; permission conveyed through Copyright Clearance Center, Inc.

can be treated independently in uncoupled

responsive to the surrounding fluctuations


negligible influence on gas phase; phases

Turbulent frozen limit. Particles are not


Mean flow frozen limit. Particles have

approached in any region of a rocket


and do not experience any turbulent

Not applicable. This limit is never


diffusion.

exhaust.
fashion.
1

than gas phase as characterized by their properties, the

Fully coupled solution. A two-way interactive coupling


Mean flow nonequilibrium. Particles cannot track gas

Turbulent nonequilibrium. Particles respond partially


with some degree of lag. They will diffuse differently

local mean flow properties and the local turbulence


and the two phases have different properties;
nonequilibrium approach utilized.

between the phases is required.

Fig. 23. Distribution of gas temperature and particle temperature on the axis
for a particle-laden jet. Reproduced with permission of AMERICAN INSTITUTE
OF AERONAUTICS AND ASTRONAUTICS, INC, from Ref. [181]; permission
conveyed through Copyright Clearance Center, Inc.
properties.

[182]. The third parameter is the mass fraction ratio, which is similar to
particle loading. The interaction induced by the particles are more
1

significant with a higher particle loading.


The interaction between the particles and the mean flow is affected
Mean flow equilibrium limit. Particles follow

respond fully to the surrounding fluctuations

by the characteristic length scales of the particles and the mean flow.
the motion of the gas phase and have the

neglected; one-way coupling utilized for


same properties; equilibrated two-phase

loading; influence on gas phase can be

Dash et al. [181] have simulated the rocket exhaust laden with alumina
One-way coupling. Negligible particle
Turbulent equilibrium limit. Particles

particles of varying sizes, under fixed nozzle operational conditions and


and diffuse like gas phase species.

a fixed particle loading. As shown in Fig. 23, the mean flow frozen limit
is attained when the particle size approaches infinity, while the mean
flow equilibrium limit is attained, when the particle size approaches 0.
The influence of the particle size on the mean flow is bounded by the
approach utilized.

two limits. For smaller particles, the exchange of momentum and en-
particle solution

ergy between the two phases is more sufficient, the change of gas ve-
locity and temperature on the jet axis is slower, and the difference
between the temperature distributions of the two phases is smaller.
1

Also, the Mach disk moves downstream and the angle between slip-
streams downstream of the Mach disk expands.
flow and characterizes the magnitude of mean flow lag

Characterizes the response of partiles to the gas phase

Characterizes the influence the particles have on the


Indicates the ability of particles to “track” the mean

As to the transient interaction between the particles and the tur-


(i.e., differences between particle and gas phase

turbulent fluctuations and hence their ability to

bulent structures, the characteristic time scales of the particles and the
turbulence are relevant. Sinha et al. [183] have adopted a LES method
in conjunction with a Lagrangian method to simulate the dispersion of
gas phase mean flow and turbulence.

the alumina particles of three sizes in a supersonic free jet. Similar to


the behavior of particles in a subsonic free jet [184], when the St is far
less than 1, the particles and the gas phase attain kinetic equilibrium;
velocities and temperatures).

when the St is far larger than 1, particles are almost irrelevant to the
turbulent motion of gas phase; when the St is around 1, particles are
associated with the turbulent structures, but the particles could leave
the structures later.
Significance

The first two parameters, relaxation length/time, describe the re-


diffuse.

sponses of the particles to the gas phase, while the particle loading
describes the influences of the particles on the gas phase. When the
particle loading is larger than 5%, the nearfield infra-red signature of a
t * = tp/T particle relaxation
l* = lp/L particle relaxation

rocket exhaust could be strongly dependent on the particles [185].


time/turbulent time

When the particle loading smaller than 0.3%, though large particles
particle mass
length/mean flow

fraction/gas mass

show great nonequilibrium, the influence on the gas temperature on the


jet axis is negligible [186]. When the alumina mass fraction at the
length scale

nozzle inlet is above 30%, the maximum gas temperature increase on


fraction
Parameter

scale

the jet axis could reach 3000K [187]. The statistics of the particle-laden
p/
Table 1

free jet experiments [182,188,189] and computations [190,191] all


* =

indicate that the Mach disk moves upstream when the particle loading

17
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

increases. In addition, smaller particles are easier to disperse and have existing measurement data, so that some critical nondimensional
greater influence on shock structures under a fixed particle loading parameters might be extracted to correlate the sub-scale experiment
[182]. and the full-scale launch.
The similarity of the flow in the free jet region between particle-free For the impinging jet interacting with launch structures, noise
impinging jets and free jets also works for particle-laden impinging jets sources under varying lift-off heights have been identified. The launch
and free jets. In another word, at a given particle size and particle platform and the deflector system are confirmed to be the two main
loading, the behavior of the particles in the free jet region of impinging noise sources during the rocket launching. It is readily acknowledged
jets is relatively consistent with that in free jets. The Direct Numerical that the launch platform would generate noise by the impingement of
Simulation (DNS) of a particle-laden impinging jet [59,192] showed the engine exhaust, in combination with the exertion of shelter effects
that small particles follow the turbulent motion as that in a free jet. on the noise originated below as well as the reflection effects on the jet
When the St of the particles is near 1, the impact velocity of particles in noise. However, the launch platform is usually studied as a part of a
an under-expanded jet is higher than that in an over-expanded jet launch complex, for which the evaluation of the specific effects is dif-
[193]. Considering the influence of the near-wall flow structures on ficult to attain. The impact of the impinging jet on single-sided de-
particles, significant deceleration of the particles is observed when the flectors, ranging from a curved plate to the deflector extended by a
particles cross the standoff shock, and the deceleration is more no- trench/duct, has been widely studied. It is found that the acoustic
ticeable for smaller particles [194]. Also, the recirculation region has sources for the engine exhaust impinging on single-sided deflectors are
significant influence on the movement of particles and thus the particle similar to the ones induced by an inclined impinging jet on a plate. For
impact on impinged objects [58]. open deflectors, the shape of the deflector is of vital importance. The
The discrete particles exchange momentum and energy with the findings of the parametric studies on the inclined impinging jet could be
surrounding flows. The influence of the jet on the particles has been partially extended to the design of the deflector surface. However, for
extensive studied, as indicated in the review of computations and ex- closed deflectors, the shape of the duct exit could be an influential
periments on cold spray [195], which emphases on the characteristics factor.
of particles in impinging jets. However, the influence of the particles on In the future work concerning the impinging jet interacting with
the jet are less studied, although it has been indicated that, under launch structures, it is recommended to study each relevant structure
certain conditions, this influence should not be neglected. Under separately at first to acquire the individual effects before considering
varying pressure ratios, impingement distances and particle sizes, the the combined effects of the complete system. The jet normally im-
particles could restrain the acceleration of the gas and weaken the pinging on a plate with cut-outs might be a reasonable model for si-
shock intensity, then the Mach disk and the standoff shock would move mulating the engine exhaust impinging on the launch platform. In ad-
upstream, and the impingement region would expand [57]. In addition, dition, multiple-sided deflectors are supposed to have lower acoustic
the particles of high temperature could exert a significant influence on efficiency than single-sided deflectors. The problem remains less stu-
the wall temperature distribution [56]. died presently. Still, multiple-sided deflectors could be applied to the
rocket producing multiple streams of engine exhaust during launching,
5. Conclusions and futrue work which ia a situation that needs to be further investigated.
The multiphase flows involved in the impingement of engine ex-
Several issues involved in impinging jets during rocket launching haust during the rocket launching include aqueous jets from water in-
are reviewed in this paper. Extensive results for two simplified models, jection system and discrete particles produced by combustion of solid
i.e. normal and inclined jets impinging on flat plates, are firstly re- propellants. The water injection system is mostly studied in sub-scale
viewed. The interactive mechanisms between impinging jets and the model experiments, where the reduction of heat and noise under dif-
launch platform or the deflector system are subsequently discussed, ferent operational conditions of the aqueous jets is qualitatively ana-
together with relevant factors that affect the flowfield and noise. lysed. Optimum heat reduction is usually attained when the water is
Attentions are then paid to the interaction of impinging jets with injected near the structures, while optimum noise reduction requires
multiphase flows, with special emphases on the issues of noise sup- water injection near the noise sources, which vary with the lift-off
pression by water injections and flow characteristics affected by dis- heights. The behavior of discrete particles in a jet has been widely
crete particles in the jet exhaust. studied and the interaction between particles and surrounding flow is
The interaction of an impinging jet with a plate has been extensively still qualitatively recognized. At the present stage, the influences of
studied. It appears that the interactive flow structures are qualitatively particles on the surrounding flow receives less emphasised. Only a few
understood and their connection with three main types of impinging studies are found concerning the particle-laden impinging jets for
noise are identified. Regardless the nearfield and farfield spectra are rocket launching.
readily measured at all relevant directions, which enables the re- Numerical investigations related to the water injection system
construction of the acoustic field as well as the identification of the during rocket launching are also encouraged, to provide more details
noise sources and the propagation of the acoustic waves, the directivity for quantitative analyses and the optimum design of the water injection
and the spectra contributed by an individual source receives less at- system. For particle-laden impinging jets, the influence of particles on
tention. Moreover, the relationships between the impinging noise and the flow structures should be emphasised. In addition, it is re-
particular flow structures, such as the recirculation region, remains commended to have a test for the influences of discrete particles on the
unclear. The qualitative trends concerning the impact on flow struc- acoustic characteristics of the impinging jet.
tures and acoustic characteristics by the relevant factors, such as the jet
Mach number, pressure ratio, and impingement distance, remain to be
straightened. Still, the thermal influences of the impinging jet are re- Acknowledgement
latively less studied. It is also difficult to predict whether and which
unstable oscillation mode would appear, so is the prediction of transi- This work was supported by grants from the National Natural
tion between different modes. Science Foundation of China (No. 11372028, 11721202), the Aerospace
Extensive analyses on the relevant factors are proposed for the Science & Technology Joint Foundation of China Aerospace Science &
purpose of fundamental studies in the future, to understand the aero- Technology Corporation (No. 6141B06220405), and a key project of
thermodynamics of the impinging jet, in particular, the high-tempera- the Science & Technology on Reliability & Environmental Engineering
ture effects on the launch site. It is likely to facilitate the understanding Laboratory (No. 614200403020517).
of the acoustic characteristics induced by different sources better with

18
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

References afterburning of rocket motor exhaust plume, Chin. J. Aeronaut. 23 (6) (2010)
653–659.
[31] K.M. Eldred, Acoustic loads generated by the propulsion system, Tech. Rep. (1971)
[1] J.K. Ignatius, S. Sathiyavageeswaran, S.R. Chakravarthy, Hot-flow simulation of SP-8072, NASA-SP-8072.
aeroacoustics and suppression by water injection during rocket liftoff, AIAA J. 53 [32] J. Varnier, W. Raguenet, D. Gely, Noise radiated from free and impinging hot
(1) (2015) 235–245. supersonic jets, 4th AIAA/CEAS Aeroacoustics Conference, American Institute of
[2] N. Karthikeyan, L. Venkatakrishnan, Flow field and acoustic investigations of the Aeronautics and Astronautics, Toulouse, France, 1998.
launch vehicle environment during lift-off, 21st AIAA/CEAS Aeroacoustics [33] J. Varnier, W. Raguenet, Experimental characterization of the sound power ra-
Conference, American Institute of Aeronautics and Astronautics, Dallas, Texas, diated by impinging supersonic jets, AIAA J. 40 (5) (2002) 825–831.
2015. [34] S. Alestra, I. Terrasse, B. Troclet, Inverse method for identification of acoustic
[3] N. Karthikeyan, L. Venkatakrishnan, Acoustic characterization of jet interaction sources at launch vehicle liftoff, AIAA J. 41 (10) (2003) 1980–1987.
with launch structures during lift-off, J. Spacecr. Rocket. 54 (2) (2017) 356–367. [35] D. Casalino, M. Barbarino, M. Genito, V. Ferrara, Hybrid empirical/computational
[4] L.M. Myers, N. Rudenko, D.K. McLaughlin, Investigation on the flow-field of two aeroacoustics methodology for rocket noise modeling, AIAA J. 47 (6) (2009)
parallel round jets impinging normal to a flat surface, 54th AIAA Aerospace 1445–1460.
Sciences Meeting, American Institute of Aeronautics and Astronautics, 2016. [36] D. Casalino, S. Santini, M. Genito, V. Ferrara, Rocket noise sources localization
[5] M. Harmon, V.N. Bhargav, P. Sellappan, F.S. Alvi, R. Kumar, Experimental study of through a CAA-based beam-forming technique, 17th AIAA/CEAS Aeroacoustics
impinging jet flow field involving Converging and CD nozzle pair, AIAA Aerospace Conference (32nd AIAA Aeroacoustics Conference), American Institute of
Sciences Meeting, American Institute of Aeronautics and Astronautics, Kissimmee, Aeronautics and Astronautics, Portland, Oregon, 2011.
Florida, 2018. [37] D. Casalino, S. Santini, M. Genito, V. Ferrara, Rocket noise sources localization
[6] M. Zhang, G. Cai, Z. Tang, B. He, W. Zhang, Y. Shu, Experimental and numerical through a tailored beam-forming technique, AIAA J. 50 (10) (2012) 2146–2158.
research on the diversion effect of a conic flame deflector for a lunar module [38] M.M.M. Morshed, Investigation of External Acoustic Loadings on a Launch Vehicle
ascent stage, J. Aerosp. Eng. 29 (5) (2016) 04016021. Fairing during Lift-Off, Ph.D. thesis University of Adelaide, 2008.
[7] A. Kagenov, A. Glazunov, K. Kostyushin, I. Eremin, V. Shuvarikov, Numerical [39] M.M.M. Morshed, C.H. Hansen, A.C. Zander, Prediction of acoustic loads on a
investigation of the effect of the configuration of ExoMars landing platform pro- launch vehicle fairing during liftoff, J. Spacecr. Rocket. 50 (1) (2013) 159–168.
pulsion system on the interaction of supersonic jets with the surface of Mars, AIP [40] M.M.M. Morshed, C.H. Hansen, A.C. Zander, Prediction of acoustic loads on a
Conference Proceedings 1893, vol. 1, 2017030084. launch vehicle: nonunique source allocation method, J. Spacecr. Rocket. 52 (5)
[8] T.B. Nickels, A.E. Perry, An experimental and theoretical study of the turbulent (2015) 1478–1485.
coflowing jet, J. Fluid Mech. 309 (1996) 157–182. [41] P. Malbqui, R. Davy, C. Bresson, Experimental characterization of the acoustics of
[9] A. Poubeau, R. Paoli, A. Dauptain, F. Duchaine, G. Wang, Large-Eddy simulations the future Ariane 6 launch pad, 7th European Conference for Aeronautics and
of a single-species solid rocket booster jet, AIAA J. 53 (6) (2014) 1477–1491. Space Sciences (EUCASS), Milan, Italy, 2017.
[10] Z. Gao, C. Jiang, C.-H. Lee, Representative interactive flamelet model and fla- [42] G. Dumnov, D. Mel'Nikov, V. Komarov, Acoustic loads on rockets during launch,
melet/progress variable model for supersonic combustion flows, Proc. Combust. 36th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, American
Inst. 36 (2) (2017) 2937–2946. Institute of Aeronautics and Astronautics, Huntsville, Alabama, 2000.
[11] A. Butt, C. Popp, H. Pitts, D. Sharp, NASA ares I launch vehicle roll and reaction [43] S. Tsutsumi, R. Takaki, E. Shima, K. Fujii, M. Arita, Generation and propagation of
control systems design status, 45th AIAA/ASME/SAE/ASEE Joint Propulsion pressure waves from H-IIA launch vehicle at lift-off, 46th AIAA Aerospace Sciences
Conference and Exhibit, American Institute of Aeronautics and Astronautics, Meeting and Exhibit, American Institute of Aeronautics and Astronautics, Reno,
Denver, Colorado, 2009. Nevada, 2008.
[12] A. Butt, C. Popp, K. Holt, H. Pitts, NASA ares I launch vehicle first stage roll control [44] Y. Nakai, N. Fujimatsu, K. Fujii, Experimental study of underexpanded supersonic
system cold flow development test program overview, 46th AIAA/ASME/SAE/ jet impingement on an inclined flat plate, AIAA J. 44 (11) (2006) 2691–2699.
ASEE Joint Propulsion Conference and Exhibit, American Institute of Aeronautics [45] Y. Goto, K. McIlroy, T. Nonomura, K. Fujii, Detailed analysis of flat plate pressure
and Astronautics, Nashville, Tennessee, 2010. peaks created by supersonic jet impingements, 47th AIAA Aerospace Sciences
[13] K. Mahesh, The interaction of jets with crossflow, Annu. Rev. Fluid Mech. 45 (1) Meeting Including the New Horizons Forum and Aerospace Exposition, American
(2013) 379–407. Institute of Aeronautics and Astronautics, Orlando, Florida, 2009.
[14] E. Hassan, J. Boles, H. Aono, D. Davis, W. Shyy, Supersonic jet and crossflow [46] Y. Nagata, T. Nonomura, K. Fujii, M. Yamamoto, Analysis of acoustic-fields gen-
interaction: computational modeling, Prog. Aero. Sci. 57 (2013) 1–24. erated by supersonic jet impinging on an inclined flat plate and a curved plate,
[15] Z. Wang, C. Jiang, Z. Gao, C. Lee, Prediction for the separation length of two- APCOM and ISCM, Singapore, 2013.
dimensional sonic injection with high-speed crossflow, AIAA J. (2017) 1–16. [47] S. Salehian, K. Kourbatski, V.V. Golubev, R.R. Mankbadi, Numerical aspects of
[16] P. Moraes, A. Morgenstern, Evaluation of the effectiveness of VLS launcher flame rocket lift-off noise with launch-pad aqueous injection, AIAA Aerospace Sciences
deflector, Comput. Fluid 30 (5) (2001) 523–532. Meeting, American Institute of Aeronautics and Astronautics, Kissimmee, Florida,
[17] Soyuz-2. RIANovosti, 1a launch vehicle carrying three Russian spacecraft mikhail 2018.
lomonosov, aist-2d, and samsat-218 at the launch pad at Vostochny space launch [48] A. Rodionov, Y. Plastinin, J. Drakes, M. Simmons, I.R. Hiers, Modeling of multi-
Centre, RIA novosti, http://en.kremlin.ru/events/president/news/51811/photos/ phase alumina-loaded jet flow fields, 34th AIAA/ASME/SAE/ASEE Joint
44005, (2016). Propulsion Conference and Exhibit, American Institute of Aeronautics and
[18] B. Ingalls, Antares Rocket Preperation, NASA Goddard Space Flight Center, (2013) Astronautics, Cleveland, OH, 1998.
https://www.flickr.com/photos/gsfc/8654648539/. [49] F.T. Buckley, R.B. Myers, Studies of solid rocket exhaust afterburning with a finite
[19] RocketLab, Electron ’its a test’ at rocket Lab LC-1, Rocket Lab (2016), http://www. difference equilibrium chemistry plume model, JANNAF 6th Plume Technology
rocketlabusa.com/gallery/. Meeting, 1971.
[20] J.K. Ignatius, S. Sankaran, R.A. Kumar, T.N.V. Satyanarayana, S.R. Chakravarthy, [50] J.W. Calhoon, Evaluation of afterburning cessation mechanisms in fuel rich rocket
Suppression of jet noise by staged water injection during launch vehicle lift-off, exhaust plumes, 34th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and
Int. J. Aeroacoustics 7 (3–4) (2008) 223–241. Exhibit, American Institute of Aeronautics and Astronautics, 1998.
[21] R. Garcia, Assessment of microphone phased array for measuring launch vehicle [51] D. Gely, G. Elias, C. Bresson, H. Foulon, S. Radulovic, P. Roux, Reduction of su-
lift-off acoustics, Tech. Rep. (2012) NASA/TM-2012-217563, NASA Langley personic jet noise - application to the Ariane 5 launch vehicle, 6th AIAA/CEAS
Research Center. Aeroacoustics Conference, American Institute of Aeronautics and Astronautics,
[22] J. Panda, R. Mosher, Microphone phased array to identify liftoff noise sources in Lahaina, Hawaii, 2000.
model-scale tests, J. Spacecr. Rocket. 50 (5) (2013) 1002–1012. [52] S. Hu, C. Jiang, Z. Gao, C.-H. Lee, Disturbance region update method for steady
[23] J. Panda, R.N. Mosher, B.J. Porter, Noise source identification during rocket en- compressible flows, Comput. Phys. Commun. 229 (2018) 68–86.
gine test firings and a rocket launch, J. Spacecr. Rocket. 51 (6) (2014) 1761–1772. [53] J. Sachdev, V. Ahuja, A. Hosangadi, D. Allgood, Analysis of flame deflector spray
[24] B. Henderson, The connection between sound production and jet structure of the nozzles in rocket engine test stands, 46th AIAA/ASME/SAE/ASEE Joint Propulsion
supersonic impinging jet, J. Acoust. Soc. Am. 111 (2) (2002) 735–747. Conference and Exhibit, American Institute of Aeronautics and Astronautics,
[25] Mcilroy2007 K. McIlroy, K. Fujii, Computational analysis of supersonic under- Nashville, Tennessee, 2010.
expanded jets impinging on an inclined flat plate, 37th AIAA Fluid Dynamics [54] P. Ragaller, A. Annaswamy, J. Gustavsson, F. Alvi, Impinging jet noise suppression
Conference and Exhibit, American Institute of Aeronautics and Astronautics, using water microjets, 49th AIAA Aerospace Sciences Meeting Including the New
Miami, Florida, 2007. Horizons Forum and Aerospace Exposition, American Institute of Aeronautics and
[26] R. Vinze, S. Chandel, M.D. Limaye, S.V. Prabhu, Heat transfer distribution and Astronautics, Orlando, Florida, 2011.
shadowgraph study for impinging underexpanded jets, Appl. Therm. Eng. 115 [55] S. Anant, T.J. Tharakan, R.P. Thomas, CFD analysis of water cooled flame de-
(2017) 41–52. flector in rocket engine test facility, Proceedings of the 5th International and 41st
[27] A. Powell, The sound-producing oscillations of round underexpanded jets im- National Conference on FMFP 2014, Springer India, New Delhi, 2017, pp.
pinging on normal plates, J. Acoust. Soc. Am. 83 (2) (1988) 515–533. 517–528.
[28] R. Reed, V. Calia, Review of aluminum oxide rocket exhaust particles, 28th [56] N. Tsuboi, A.K. Hayashi, T. Fujiwara, K. Arashi, M. Kodama, Numerical simulation
Thermophysics Conference, American Institute of Aeronautics and Astronautics, of a supersonic jet impingement on a ground, SAE Trans. 100 (1) (1991)
Orlando, Florida, 1993. 2168–2180.
[29] J. Troyes, I. Dubois, V. Borie, A. Boischot, Multi-phase reactive numerical simu- [57] G. Wang, W. Xu, Y. Gao, A study on erosive effect of two phase impinging jets on
lations of a model solid rocket exhaust jet, 42nd AIAA/ASME/SAE/ASEE Joint vertical plate, J. Ballist. 21 (02) (2009) 10–14.
Propulsion Conference and Exhibit, American Institute of Aeronautics and [58] N.A. Buchmann, C. Cierpka, C.J. Khler, J. Soria, Ultra-high-speed 3D astigmatic
Astronautics, Sacramento, California, 2006. particle tracking velocimetry: application to particle-laden supersonic impinging
[30] J. Yi, M. Yanli, W. Weichen, S. Liwu, Inhibition effect of water injection on jets, Exp. Fluid 55 (11) (2014) 1842.

19
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

[59] K. Akhtar, S.A. Ragab, Numerical simulations of a particle-laden rectangular su- [94] T.B. Davis, A. Edstrand, L.N. Cattafesta, F.S. Alvi, D. Yorita, K. Asai, Investigation
personic jet impinging on a solid wall, 52nd Aerospace Sciences Meeting, of the instabilities of supersonic impinging jets using unsteady pressure sensitive
American Institute of Aeronautics and Astronautics, National Harbor, Maryland, paint, 52nd Aerospace Sciences Meeting, American Institute of Aeronautics and
2014. Astronautics, National Harbor, Maryland, 2014.
[60] C.D. Donaldson, R.S. Snedeker, A study of free jet impingement. Part 1. Mean [95] T. Davis, A. Edstrand, F. Alvi, L. Cattafesta, D. Yorita, K. Asai, Investigation of
properties of free and impinging jets, J. Fluid Mech. 45 (2) (1971) 281–319. impinging jet resonant modes using unsteady pressure-sensitive paint measure-
[61] C.K.W. Tam, Supersonic jet noise, Annu. Rev. Fluid Mech. 27 (1) (1995) 17–43. ments, Exp. Fluid 56 (5) (2015) 101.
[62] E. Franquet, V. Perrier, S. Gibout, P. Bruel, Free underexpanded jets in a quiescent [96] J. Panda, Shock oscillation in underexpanded screeching jets, J. Fluid Mech. 363
medium: a review, Prog. Aero. Sci. 77 (2015) 25–53. (1998) 173–198.
[63] A. Krothapalli, E. Rajkuperan, F. Alvi, L. Lourenco, Flow field and noise char- [97] P.C. Stegeman, J.M. Perez, J. Soria, V. Theofilis, Inception and evolution of co-
acteristics of a supersonic impinging jet, J. Fluid Mech. 392 (1999) 155–181. herent structures in under-expanded supersonic jets, J. Phys. Conf. Ser. 708 (1)
[64] R. Vinze, M.D. Limeye, S.V. Prabhu, Influence of the elliptical and circular orifices (2016) 012015.
on the local heat transfer distribution of a flat plate impinged by under-expanded [98] P.J. Schmid, Dynamic mode decomposition of numerical and experimental data, J.
jets, Heat Mass Transf. 53 (4) (2017) 1439–1455. Fluid Mech. 656 (2010) 5–28.
[65] G. Sinibaldi, L. Marino, G.P. Romano, Sound source mechanisms in under-ex- [99] K. Shahram, P. C, Linearised dynamics and non-modal instability analysis of an
panded impinging jets, Exp. Fluid 56 (5) (2015) 105. impinging under-expanded supersonic jet, J. Phys. Conf. Ser. 1001 (1) (2018)
[66] M.V. Silnikov, M.V. Chernyshov, Incident shock strength evolution in over- 012019.
expanded jet flow out of rocket nozzle, Acta Astronaut. 135 (2017) 172–180. [100] G. Sinibaldi, G. Lacagnina, L. Marino, G.P. Romano, Analysis of the characteristic
[67] R. Gojon, C. Bogey, O. Marsden, Investigation of tone generation in ideally ex- acoustic tones of an impinging jet, 19th AIAA/CEAS Aeroacoustics Conference,
panded supersonic planar impinging jets using large-eddy simulation, J. Fluid American Institute of Aeronautics and Astronautics, 2013.
Mech. 808 (2016) 90–115. [101] J. Gustavsson, P. Ragaller, R. Kumar, F. Alvi, Temperature effect on acoustics of
[68] R. Gojon, C. Bogey, Flow features near plate impinged by ideally expanded and supersonic impinging jet, 16th AIAA/CEAS Aeroacoustics Conference, American
underexpanded round jets, AIAA J. 56 (2) (2017) 445–457. Institute of Aeronautics and Astronautics, 2010.
[69] A. Dauptain, B. Cuenot, L.Y.M. Gicquel, Large eddy simulation of stable supersonic [102] D.M. Mitchell, D.R. Honnery, J. Soria, The visualization of the acoustic feedback
jet impinging on flat plate, AIAA J. 48 (10) (2010) 2325–2338. loop in impinging underexpanded supersonic jet flows using ultra-high frame rate
[70] A. Dauptain, L.Y.M. Gicquel, S. Moreau, Large eddy simulation of supersonic schlieren, J. Vis. 15 (4) (2012) 333–341.
impinging jets, AIAA J. 50 (7) (2012) 1560–1574. [103] J. L. Weightman, O. Amili, D. Honnery, J. Soria, D. Edgington-Mitchell, An ex-
[71] J. Iwamoto, Impingement of under-expanded jets on a flat plate, J. Fluids Eng. 112 planation for the phase lag in supersonic jet impingement, J. Fluid Mech. 815.
(2) (1990) 179–184. [104] B. Henderson, J. Bridges, M. Wernet, An experimental study of the oscillatory flow
[72] J.C. Carling, B.L. Hunt, The near wall jet of a normally impinging, uniform, ax- structure of tone-producing supersonic impinging jets, J. Fluid Mech. 542 (2005)
isymmetric, supersonic jet, J. Fluid Mech. 66 (1) (1974) 159–176. 115–137.
[73] C. Chin, M. Li, C. Harkin, T. Rochwerger, L. Chan, A. Ooi, A. Risborg, J. Soria, [105] T. Worden, J. Gustavsson, C. Shih, F.S. Alvi, Acoustic measurements of high-
Investigation of the flow structures in supersonic free and impinging jet flows, J. temperature supersonic impinging jets in multiple configurations, 19th AIAA/
Fluids Eng. 135 (3) (2013) 031202–031202–12. CEAS Aeroacoustics Conference, American Institute of Aeronautics and
[74] L. Chan, C. Chin, J. Soria, A. Ooi, Large eddy simulation and Reynolds-averaged Astronautics, Berlin, Germany, 2013.
Navier-Stokes calculations of supersonic impinging jets at varying nozzle-to-wall [106] C. Brehm, J.A. Housman, C.C. Kiris, Noise generation mechanisms for a supersonic
distances and impinging angles, Int. J. Heat Fluid Flow 47 (2014) 31–41. jet impinging on an inclined plate, J. Fluid Mech. 797 (2016) 802–850.
[75] S.G. Kundasev, N.P. Kiselev, V.I. Zapryagaev, Experimental investigation of the [107] M. Akamine, Y. Nakanishi, K. Okamoto, S. Teramoto, T. Okunuki, S. Tsutsumi,
flow structure of the supersonic jet impinging on an inclined flat obstacle, AIP Experimental study on acoustic phenomena of supersonic jet impinging on in-
Conference Proceedings 1770 (1) (2016) 030031. clined flat plate, 52nd Aerospace Sciences Meeting, American Institute of
[76] K. Fujii, A. Oyama, N. Tsuboi, M. Tsukada, H. Ouchi, M. Ito, K. Hayashi, Flow field Aeronautics and Astronautics, National Harbor, Maryland, 2014.
analysis of under-expanded supersonic jets impinging on an inclined flat plate: [108] M. Akamine, Y. Nakanishi, K. Okamoto, S. Teramoto, T. Okunuki, S. Tsutsumi,
analysis with PSP/schlieren images and CFD simulations, 2005 ASME Fluids Acoustic phenomena from correctly expanded supersonic jet impinging on in-
Engineering Division Summer Meeting and Exhibitio, Houston, Texas, 2005, pp. clined plate, AIAA J. 53 (7) (2015) 2061–2067.
213–221. [109] M. Akamine, K. Okamoto, S. Teramoto, T. Okunuki, S. Tsutsumi, Conditional
[77] J. Seiner, Advances in high speed jet aeroacoustics, AIAA/NASA 9th Aeroacoustics sampling analysis of acoustic phenomena from a supersonic jet impinging on an
Conference, American Institute of Aeronautics and Astronautics, Hampton, inclined flat plate, Trans. Jpn. Soc. Aeronaut. Space Sci. 59 (5) (2016) 287–294.
Virginia, 1984. [110] M. Akamine, K. Okamoto, K.L. Gee, T.B. Neilsen, S. Teramoto, T. Okunuki,
[78] G. Raman, Advances in understanding supersonic jet screech, 36th Aerospace S. Tsutsumi, Effect of nozzle-plate distance on acoustic phenomena from super-
Sciences Meeting and Exhibit, American Institute of Aeronautics and Astronautics, sonic impinging jet, 22nd AIAA/CEAS Aeroacoustics Conference, American
Reno, Nevada, 1998. Institute of Aeronautics and Astronautics, Lyon, France, 2016.
[79] G. Raman, ADVANCES IN UNDERSTANDING SUPERSONIC JET SCREECH: [111] T. Nonomura, Y. Goto, K. Fujii, Acoustic waves from a supersonic jet impinging on
REVIEW AND PERSPECTIVE, Prog. Aero. Sci. 34 (1) (1998) 45–106. an inclined flat plate, 48th AIAA Aerospace Sciences Meeting Including the New
[80] G. Raman, SUPERSONIC JET SCREECH: HALF-CENTURY FROM POWELL TO THE Horizons Forum and Aerospace Exposition, American Institute of Aeronautics and
PRESENT, J. Sound Vib. 225 (3) (1999) 543–571. Astronautics, Orlando, Florida, 2010.
[81] J.C. Yu, D.S. Dosanjh, Noise field of a supersonic mach 1.5 cold model jet, J. [112] T. Nonomura, Y. Goto, K. Fujii, Aeroacoustic waves generated from a supersonic
Acoust. Soc. Am. 51 (5A) (1972) 1400–1410. jet impinging on an inclined flat plate, Int. J. Aeroacoustics 10 (4) (2011)
[82] T.D. Norum, J.M. Seiner, Measurements of mean static pressure and far field 401–425.
acoustics of shock containing supersonic jets, Tech. Rep. (1982) NASA-TM-84521. [113] T. Nonomura, K. Fujii, POD of aeroacoustic fields of a jet impinging on an inclined
[83] H.K. Tanna, An experimental study of jet noise part I: turbulent mixing noise, J. plate, 16th AIAA/CEAS Aeroacoustics Conference, American Institute of
Sound Vib. 50 (3) (1977) 405–428. Aeronautics and Astronautics, 2010.
[84] J. Seiner, T. Norum, Experiments of shock associated noise of supersonic jets, 12th [114] T. Nonomura, K. Fujii, Computational study of effects of near-wall turbulent
Fluid and Plasma Dynamics Conference, American Institute of Aeronautics and structure on aeroacoustic waves from a supersonic jet impinging on a inclined
Astronautics, Hampton, Virginia, 1979. plate, 17th AIAA/CEAS Aeroacoustics Conference (32nd AIAA Aeroacoustics
[85] J. Seiner, T. Norum, Aerodynamic aspects of shock containing jet plumes, 6th Conference), American Institute of Aeronautics and Astronautics, Portland,
Aeroacoustics Conference, American Institute of Aeronautics and Astronautics, Oregon, 2011.
Hampton, Virginia, 1980. [115] S. Tsutsumi, T. Nonomura, K. Fujii, Y. Nakanishi, K. Okamoto, S. Teramoto,
[86] J.M. Seiner, J.C. Yu, Acoustic near-field properties associated with broadband Analysis of acoustic wave from supersonic jets impinging to an inclined flat plate,
shock noise, AIAA J. 22 (9) (1984) 1207–1215. Seventh International Conference on Computational Fluid Dynamics (ICCFD7), Big
[87] T.D. Norum, J.M. Seiner, Broadband shock noise from supersonic jets, AIAA J. 20 Island, Hawaii, 2012.
(1) (1982) 68–73. [116] H. Honda, T. Nonomura, K. Fujii, M. Yamamoto, Effects of plate Angles on acoustic
[88] A. Powell, On the mechanism of choked jet noise, Proc. Phys. Soc. B 66 (12) waves from a supersonic jet impinging on an inclined flat plate, 41st AIAA Fluid
(1953) 1039. Dynamics Conference and Exhibit, American Institute of Aeronautics and
[89] A. Powell, The noise of choked jets, J. Acoust. Soc. Am. 25 (3) (1953) 385–389. Astronautics, Honolulu, Hawaii, 2011.
[90] J. Seiner, J. Manning, M. Ponton, Model and full scale study of twin supersonic [117] T. Nonomura, H. Honda, Y. Nagata, M. Yamamoto, S. Morizawa, S. Obayashi,
plume resonance, 25th Aerospace Sciences Meeting, American Institute of K. Fujii, Plate-angle effects on acoustic waves from supersonic jets impinging on
Aeronautics and Astronautics, Reno, Nevada, 1987. inclined plates, AIAA J. 54 (3) (2016) 816–827.
[91] B. Henderson, A. Powell, Experiments concerning tones produced by an axisym- [118] G. Berkooz, P. Holmes, J.L. Lumley, The proper orthogonal decomposition in the
metric choked jet impinging on flat plates, J. Sound Vib. 168 (2) (1993) 307–326. analysis of turbulent flows, Annu. Rev. Fluid Mech. 25 (1) (1993) 539–575.
[92] B. Henderson, An experimental investigation into the sound producing char- [119] J. Panda, R.G. Seasholtz, Experimental investigation of density fluctuations in
acteristics of supersonic impinging jets, 7th AIAA/CEAS Aeroacoustics Conference, high-speed jets and correlation with generated noise, J. Fluid Mech. 450 (2002)
American Institute of Aeronautics and Astronautics, Maastricht, Netherlands, 97–130.
2001. [120] J. Panda, Identification of noise sources in high speed jets via correlation mea-
[93] A. Risborg, J. Soria, High-speed optical measurements of an underexpanded su- surements - a review, 11th AIAA/CEAS Aeroacoustics Conference (26th AIAA
personic jet impinging on an inclined plate, 28th International Congress on High- Aeroacoustics Conference), American Institute of Aeronautics and Astronautics,
Speed Imaging and Photonics, 2009 71261F–71261F–11. Monterey, California, 2005.

20
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

[121] C. Bogey, C. Bailly, An analysis of the correlations between the turbulent flow and axisymmetric deflector, AIAA J. 32 (7) (1994) 1535–1538.
the sound pressure fields of subsonic jets, J. Fluid Mech. 583 (2007) 71–97. [150] P.J. Lamont, B.L. Hunt, The impingement of underexpanded axisymmetric jets on
[122] M. Akamine, K. Okamoto, K.L. Gee, T.B. Neilsen, S. Teramoto, T. Okunuki, wedges, J. Fluid Mech. 76 (2) (1976) 307–336.
S. Tsutsumi, Effect of NozzlePlate distance on acoustic phenomena from super- [151] I.K. Jennions, B.L. Hunt, The axisymmetric impingement of supersonic air jets on
sonic impinging jet, AIAA J. 56 (5) (2018) 1943–1952. cones, Aeronaut. Q. 31 (1) (1980) 26–41.
[123] S. Kawai, S. Tsutsumi, R. Takaki, K. Fujii, Computational aeroacoustic analysis of [152] B.L. Hunt, Loads and pressures due to underexpanded jets impinging on wedges,
overexpanded supersonic jet impingement on a flat plate with/without hole, 5th Aeronaut. Q. 34 (2) (1983) 76–98.
Joint ASME/JSME Fluids Engineering Conference, San Diego, California USA, [153] R. Vinze, A. Chollackal, M.D. Limaye, S.V. Prabhu, Heat transfer characteristics of
2007, pp. 1163–1167. the jet deflector impinged by underexpanded jets, Exp. Therm. Fluid Sci. 78 (2016)
[124] A. Wiley, I. Choutapalli, R. Kumar, F. Alvi, Noise and flowfield characteristics of a 167–181.
supersonic jet impinging on a porous surface, 48th AIAA Aerospace Sciences [154] S.K. Patel, J. Mathew, Acoustic fields of a supersonic jet deflected by wedges
Meeting Including the New Horizons Forum and Aerospace Exposition, American mounted on a flat plate, 22nd AIAA/CEAS Aeroacoustics Conference, American
Institute of Aeronautics and Astronautics, Orlando, Florida, 2010. Institute of Aeronautics and Astronautics, Lyon, France, 2016.
[125] A. Wiley, R. Kumar, Supersonic impinging jet noise reduction using a hybrid [155] J. Housman, M. Barad, C. Kiris, Space-time accuracy assessment of CFD simula-
control technique, J. Sound Vib. 348 (2015) 88–104. tions for the launch environment, 29th AIAA Applied Aerodynamics Conference,
[126] A. Dhamanekar, K. Srinivasan, Effect of impingement surface roughness on the American Institute of Aeronautics and Astronautics, Honolulu, Hawaii, 2011.
noise from impinging jets, Phys. Fluids 26 (3) (2014) 036101. [156] H. Cai, W. Nie, K. Guo, S. Zhou, Influence of deflector on impact properties of
[127] R. Gaeta, K. Ahuja, B. Murdock, R. Combier, Noise reduction from a distributed multi-nozzle LOX/kerosene engine exhaust plume, 8th International Conference
exhaust nozzle with forward velocity effects, 10th AIAA/CEAS Aeroacoustic on Mechanical and Aerospace Engineering, ICMAE), 2017, pp. 296–300.
Conference, American Institute of Aeronautics and Astronautics, Manchester, [157] D.A. Drew, Mathematical modeling of two-phase flow, Annu. Rev. Fluid Mech. 15
England, 2004. (1) (1983) 261–291.
[128] K.K. Ahuja, D.A. Alvord, J. Mattingly, J. Mittelman, W. Schuttler, D. Dickey, [158] J.S. Shirolkar, C.F.M. Coimbra, M. Queiroz McQuay, Fundamental aspects of
Thoughts on use of university-scale rocket models to study launch acoustics, 20th modeling turbulent particle dispersion in dilute flows, Prog. Energy Combust. Sci.
AIAA/CEAS Aeroacoustics Conference, American Institute of Aeronautics and 22 (4) (1996) 363–399.
Astronautics, Atlanta, Georgia, 2014. [159] G. Gouesbet, A. Berlemont, Eulerian and Lagrangian approaches for predicting the
[129] S. Basu, S. Saha, D. Chakraborty, Numerical simulation of missile jet deflector, J. behaviour of discrete particles in turbulent flows, Prog. Energy Combust. Sci. 25
Spacecr. Rocket. 54 (4) (2017) 930–935. (2) (1999) 133–159.
[130] T.J. Worden, C. Shih, F.S. Alvi, Supersonic jet impingement on a model-scale jet [160] E. Loth, Numerical approaches for motion of dispersed particles, droplets and
blast deflector, AIAA J. 55 (8) (2017) 2522–2536. bubbles, Prog. Energy Combust. Sci. 26 (3) (2000) 161–223.
[131] R.W. Powers, D.K. McLaughlin, P.J. Morris, Noise reduction in supersonic jets [161] F. Mashayek, R.V.R. Pandya, Analytical description of particle/droplet-laden tur-
exhausting over a simulated aircraft carrier deck, J. Aircr. 55 (1) (2017) 310–324. bulent flows, Prog. Energy Combust. Sci. 29 (4) (2003) 329–378.
[132] S. Ma, J. Tan, X. Li, J. Hao, The Effect Analysis of an Engine Jet on an Aircraft Blast [162] A. Guha, Transport and deposition of particles in turbulent and laminar flow,
Deflector, Transactions of the Institute of Measurement and Control . Annu. Rev. Fluid Mech. 40 (1) (2008) 311–341.
[133] S. Tsutsumi, K. Fukuda, R. Takaki, E. Shima, K. Fujii, K. Ui, Numerical study on [163] S. Balachandar, J.K. Eaton, Turbulent dispersed multiphase flow, Annu. Rev. Fluid
acoustic radiation for designing launch-pad of advanced solid rocket, 44th AIAA/ Mech. 42 (1) (2009) 111–133.
ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, American Institute of [164] X. Jiang, G.A. Siamas, K. Jagus, T.G. Karayiannis, Physical modelling and ad-
Aeronautics and Astronautics, Hartford, Connecticut, 2008. vanced simulations of gasliquid two-phase jet flows in atomization and sprays,
[134] T. Tatsukawa, T. Nonomura, A. Oyama, K. Fujii, Multi-objective aeroacoustic Prog. Energy Combust. Sci. 36 (2) (2010) 131–167.
design exploration of launch-pad flame deflector using large-eddy simulation, J. [165] C. Jiang, C. Xu, Z. Gao, C. Lee, Finite panel method for the simulation of wind-
Spacecr. Rocket. 53 (4) (2016) 751–758. driven rain, Build. Environ. 105 (2016) 358–368.
[135] T. Tatsukawa, Y. Nagata, T. Nonomura, A. Oyama, K. Fujii, M. Yamamoto, [166] A. Bourgine, J. Dordain, Review of acoustic studies carried out at ONERA in the
Multiobjective design exploration of an aeroacoustic design problem for rocket field of space vehicles, J. Br. Interplanet. Soc. (JBIS), 34, p.72 .
launch site with evolutionary computation and large eddy simulations, 10th AIAA [167] E. Zoppellari, D. Juve, E. Zoppellari, D. Juve, Reduction of jet noise by water
Multidisciplinary Design Optimization Conference, American Institute of injection, 3rd AIAA/CEAS Aeroacoustics Conference, American Institute of
Aeronautics and Astronautics, National Harbor, Maryland, 2014. Aeronautics and Astronautics, Atlanta, Georgia, 1997.
[136] T. Tatsukawa, T. Nonomura, A. Oyama, K. Fujii, Nozzle-to-Ground distance effect [168] E. Zoppellari, D. Juve, Reduction of hot supersonic jet noise by water injection, 4th
on nondominated solutions of multiobjective aeroacoustic flame deflector design AIAA/CEAS Aeroacoustics Conference, American Institute of Aeronautics and
problem, 21st AIAA/CEAS Aeroacoustics Conference, American Institute of Astronautics, Toulouse, France, 1998.
Aeronautics and Astronautics, Dallas, Texas, 2015. [169] M. Kandula, M. Lonergan, Effective jet properties for the estimation of turbulent
[137] D. Allgood, V. Ahuja, Computational plume modeling of conceptual ARES vehicle mixing noise reduction by water injection, 13th AIAA/CEAS Aeroacoustics
stage tests, 43rd AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Conference (28th AIAA Aeroacoustics Conference), American Institute of
American Institute of Aeronautics and Astronautics, Cincinnati, Ohio, 2007. Aeronautics and Astronautics, Rome, Italy, 2007.
[138] S. Tsutsumi, S. Kato, K. Fukuda, R. Takaki, K. Ui, Effect of deflector shape on [170] M. Kandula, Prediction of turbulent jet mixing noise reduction by water injection,
acoustic field of launch vehicle at lift-off, 47th AIAA Aerospace Sciences Meeting AIAA J. 46 (11) (2008) 2714–2722.
Including the New Horizons Forum and Aerospace Exposition, American Institute [171] D. Washington, A. Krothapalli, The role of water injection on the mixing noise
of Aeronautics and Astronautics, Orlando, Florida, 2009. supersonic jet, 4th AIAA/CEAS Aeroacoustics Conference, American Institute of
[139] S. Tsutsumi, R. Takaki, Y. Nakanishi, K. Okamoto, S. Teramoto, Acoustic gen- Aeronautics and Astronautics, Toulouse, France, 1998.
eration mechanism of a supersonic jet impinging on deflectors, 52nd Aerospace [172] A. Krothapalli, L. Venkatakrishnan, L. Lourenco, B. Greska, R. Elavarasan,
Sciences Meeting, American Institute of Aeronautics and Astronautics, National Turbulence and noise suppression of a high-speed jet by water injection, J. Fluid
Harbor, Maryland, 2014. Mech. 491 (2003) 131–159.
[140] M. Kandula, B. Vu, Scale model experiments on sound propagation from a mach [173] T. Norum, Reductions in multi-component jet noise by water injection, 10th
2.5 cold nitrogen jet flowing through a rigid-walled duct with a J-deflector, Tech. AIAA/CEAS Aeroacoustics Conference, American Institute of Aeronautics and
Rep. (2003) NASA/TM-2003-211186. Astronautics, Manchester, Great Britain, 2004.
[141] M. Kandula, Sound radiation from a supersonic jet passing through a partially [174] B. Henderson, Fifty years of fluidic injection for jet noise reduction, Int. J.
open exhaust duct, J. Vib. Acoust. 133 (6) (2011) 064503–064503–5. Aeroacoustics 9 (1–2) (2010) 91–122.
[142] S. Tsutsumi, R. Takaki, T. Hara, H. Ueda, H. Nagata, Numerical analysis of ignition [175] B.T. Vu, N. Bachchan, O. Peroomian, V. Akdag, Multiphase modeling of water
overpressure effect on H-IIB launch vehicle, J. Spacecr. Rocket. 51 (3) (2014) injection on flame deflector, 21st AIAA Computational Fluid Dynamics
893–899. Conference, American Institute of Aeronautics and Astronautics, San Diego,
[143] H. Oh, J. Lee, H. Um, H. Huh, Numerical study for flame deflector design of a California, 2013.
space launch vehicle, Adv. Space Res. 59 (7) (2017) 1833–1847. [176] B.W. Bartlett, L.J. Delaney, Effect of liquid surface tension on maximum particles
[144] NASA, ATK ALV X-1 Launch Vehicle Pathfinder, (2006) https://sites.wff.nasa.gov/ size in two-phase nozzle flow, Pyrodynamics 4 (1966) 337–341.
code840/multimedia/ATK.jpg. [177] P.T. Girata, W.K. McGregor, Particle sampling of solid rocket motor exhausts in
[145] S. Tsutsumi, T. Ishii, K. Ui, S. Tokudome, K. Wada, Study on acoustic prediction high-altitude test cells, 21st Aerospace Sciences Meeting, American Institute of
and reduction of Epsilon launch vehicle at liftoff, J. Spacecr. Rocket. 52 (2) (2015) Aeronautics and Astronautics, Reno, Nevada, 1984, pp. 293–311.
350–361. [178] R.W. Hermsen, Aluminum oxide particle size for solid rocket motor performance
[146] T. Ishii, S. Tsutsumi, K. Ui, S. Tokudome, Y. Ishii, K. Wada, S. Nakamura, Acoustic prediction, J. Spacecr. Rocket. 18 (6) (1981) 483–490.
measurement of 1:42 scale booster and launch pad, Proceedings of Meetings on [179] W.D. Brennan, D.L. Hovland, D.W. Netzer, Measured particulate behavior in a
Acoustics 18 (1) (2012) 040009. subscale solid propellant rocket motor, J. Propuls. Power 8 (5) (1992) 954–960.
[147] S. Tsutsumi, T. Ishii, K. Ui, S. Tokudome, K. Wada, Assessing prediction and re- [180] H.O. Kim, D. Laredo, D.W. Netzer, Measurement of submicrometer Al2O3 particles
duction technique of lift-off acoustics using Epsilon flight data, 53rd AIAA in plumes, Appl. Opt. 32 (33) (1993) 6834–6840.
Aerospace Sciences Meeting, American Institute of Aeronautics and Astronautics, [181] S.M. Dash, D.E. Wolf, R.A. Beddini, H.S. Pergament, Analysis of two-phase flow
Kissimmee, Florida, 2015. processes in rocket exhaust plumes, J. Spacecr. Rocket. 22 (3) (1985) 367–380.
[148] J.K. Prasad, R.C. Mehta, A.K. Sreekanth, Experimental study of overexpanded [182] M. Sommerfeld, The structure of particle-laden, underexpanded free jets, Shock
supersonic jet impingement on a double wedge deflector, Aeronaut. J. 97 (966) Waves 3 (4) (1994) 299–311.
(1968) 209–214 1993. [183] N. Sinha, D. Kenzakowski, W. Calhoon, Particulate dispersion in rocket exhaust
[149] J.K. Prasad, C. Mehta, A.K. Sreekanth, Impingement of supersonic jets on an plumes, 38th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit,

21
C. Jiang, et al. Progress in Aerospace Sciences 109 (2019) 100547

American Institute of Aeronautics and Astronautics, Indianapolis, Indiana, 2002. [190] A.D. Rychkov, Flow of a mixture of gas and solid particles in supersonic under-
[184] J. Uthuppan, S.K. Aggarwal, F. Grinstein, K. Kailasanath, Particle dispersion in a expanded jets, Fluid Dyn. 9 (2) (1974) 224–227.
transitional axisymmetric jet - a numerical simulation, AIAA J. 32 (10) (1994) [191] O. Ejtehadi, A. Rahimi, A. Karchani, R.S. Myong, Complex wave patterns in dilute
2004–2014. gasparticle flows based on a novel discontinuous Galerkin scheme, Int. J. Multiph.
[185] D. Laredo, D.W. Netzer, The dominant effect of alumina on nearfield plume ra- Flow 104 (2018) 125–151.
diation, J. Quant. Spectrosc. Radiat. Transf. 50 (5) (1993) 511–530. [192] K. Akhtar, A Numerical Study of Supersonic Rectangular Jet Impingement and
[186] R. Varun, T. Sundararajan, R. Usha, K. Srinivasan, Interaction between particle- Applications to Cold Spray Technology, Ph.D. thesis Virginia Polytechnic Institute
laden underexpanded twin supersonic jets, Proc. IME G J. Aero. Eng. 224 (9) and State University, 2014.
(2010) 1005–1025. [193] S. Li, B. Muddle, M. Jahedi, J. Soria, A numerical investigation of the cold spray
[187] Z. Li, H. Xiang, Numerical simulation of the reactive two-phase solid rocket motor process using underexpanded and overexpanded jets, J. Therm. Spray Technol. 21
exhaust plume, 2Nd International Conference on Computer Science and (1) (2012) 108–120.
Electronics Engineering(ICCSEE2013), Hangzhou, China, 2013, pp. 1508–1511. [194] B. Jodoin, Cold spray nozzle mach number limitation, J. Therm. Spray Technol. 11
[188] D.J. Carlson, C.H. Lewis, Normal shock location in underexpanded gas and gas- (4) (2002) 496–507.
particle jets, AIAA J. 2 (4) (1964) 776–777. [195] S. Yin, M. Meyer, W. Li, H. Liao, R. Lupoi, Gas flow, particle acceleration, and heat
[189] J.S. Draper, P.O. Jarvinen, Underexpanded gas-particle jets, AIAA J. 5 (4) (1967) transfer in cold spray: a review, J. Therm. Spray Technol. 25 (5) (2016) 874–896.
824–825.

22

You might also like