You are on page 1of 82

Progress in Aerospace Sciences 119 (2020) 100636

Contents lists available at ScienceDirect

Progress in Aerospace Sciences


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

Review of combustion stabilization for hypersonic airbreathing propulsion


Qili Liu a, Damiano Baccarella a, b, Tonghun Lee a, *
a
Department of Mechanical Science and Engineering, University of Illinois at Urbana-Champaign, Urbana, IL, 61801, USA
b
University of Tennessee, Knoxville, USA

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

Keywords: A review of fundamental research in combustion stabilization for hypersonic airbreathing propulsion is pre­
Hypersonic propulsion sented. Combustion in high-speed airbreathing propulsion systems demands stable flame distribution and
Combustion stabilization chemical reaction to provide reliable thrust over a wide flight envelope. Various methods have been developed to
Scramjet engine
stabilize combustion depending on the hypersonic regime. For low hypersonic conditions, combustion occurs
Diffusive combustion
Shock-induced combustion
mainly in the diffusive mode in which the gas/liquid fuels are injected into supersonic freestreams for simul­
taneous fuel-air mixing and chemical reaction. Flame stability is generally enhanced by improved mixing,
physical flameholding, and external energy addition. In higher hypersonic conditions, partially/fully premixed
combustion relying on shock induced stabilization becomes more dominant. In such cases, flame stabilization can
be achieved through alternative means such as radical generation and standing oblique detonation waves. The
review outlines both experimental and numerical research progress made towards combustion stabilization over
the entire hypersonic regime, and intended to lay the groundwork for further studies which can provide opti­
mized design guidelines for the next generation of high-speed airbreathing propulsion systems.

1. Introduction actively pursued through large scale national research programs and
international collaborations in which both success and failure have
Flame stabilization in hypersonic flows is an important scientific and paved the way for modern test flight demonstrations. The following is a
technical challenge with its main implications in supersonic and hy­ brief historical overview on high-speed airbreathing research programs,
personic airbreathing propulsion systems. Since the 1940s, much effort covering a few representative efforts that pioneered the development of
has been made in studying different types of airbreathing propulsion hypersonic technology.
systems that can provide high-speed long-range flight capabilities in The initial development of the scramjet engine occurred from 1964
high Mach number conditions [1]. In particular, airbreathing supersonic to 1975 in the Hypersonic Research Engine (HRE) Program, which
combustion ramjet (scramjet) engine is essential for developing hyper­ aimed to test a regeneratively-cooled, flight-weight scramjet based on
sonic vehicles since it provides the highest specific impulse at hyper­ the North American X-15 rocket research airplane program sponsored
sonic flight conditions by taking atmospheric air and burning fuels in by the National Aeronautics and Space Administration (NASA) of the
supersonic flows. The development of scramjet propulsion technology United States (U.S.) [1]. The expectations of the experimental scramjet
mainly focuses on (1) fast long-range aircraft for commercial trans­ engine were proven to be overly optimistic and largely illuminated
portation around the globe, (2) affordable and reusable space access critical unknowns including the need for realistic ground-test simulation
systems for space exploration, and (3) advanced weapon systems for facilities, the design complications imposed by the extraordinarily wide
global strike and reconnaissance [2]. All material presented in this re­ operating range, and the difficulty of attaining efficient mixing and
view has been compiled from publicly accessible literature. combustion in scramjet combustors [1].
From 1983 to 1995, the National Aero-Space Plane (NASP) Program
was executed to explore the feasibility of the single-stage-to-orbit
1.1. Overview of high-speed airbreathing propulsion spacecraft. The program was led by the U.S. Department of Defense
and NASA through the development of a full-scale, flight-test aircraft,
Ever since the word hypersonic was first coined by Hsue-shen Tsien in Rockwell X-30, which was planned to attain orbital velocity with critical
1946 [3,4], high-speed airbreathing propulsion research has been

* Corresponding author. Department of Mechanical Science Engineering, University of Illinois at Urbana-Champaign, 1206 W. Green St., Urbana, IL, 61801, USA.
E-mail addresses: qililiu@illinois.edu (Q. Liu), baccare2@illinois.edu (D. Baccarella), tonghun@illinois.edu, tonghun@illinois.edu (T. Lee).

https://doi.org/10.1016/j.paerosci.2020.100636
Received 30 March 2020; Received in revised form 13 June 2020; Accepted 18 June 2020
Available online 20 November 2020
0376-0421/© 2020 Elsevier Ltd. All rights reserved.
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

hypersonic technologies [5]. Although the NASP program developed a regime with Mach numbers 4–8 [38,39]. These programs provide the
significant number of technical breakthroughs including the materials opportunities for industries and academic institutions across Europe to
and design methods for scramjet engines, it was an overly ambitious goal practice the indispensable cooperation needed to transition basic
that was both technically and programmatically unachievable given the high-speed airbreathing propulsion research into a practical system. The
relative immaturity of the various hypersonic propulsion technologies LAPCAT has shown that hypersonic speed transport can be achieved at
and the budgetary constraints of the time [6,7]. During the years Mach 4–5 and Mach 8 based on hydrogen fuel with a long-haul range by
1991–1998, the Russian Central Institute of Aviation Motors (CIAM) and assessing advanced propulsion concepts and related technology on
U.S. NASA performed joint efforts on rocket-boosted flight tests of a aircraft performance, including aerodynamic and propulsive achievable
combined ramjet and scramjet (dual-mode) airbreathing engine to efficiencies. Experimental tests have been conducted to verify and
provide flight demonstration of supersonic combustion, which gener­ validate performance of engine components, which include the intakes
ated data for ground-to-flight comparison for methodology verification [40–42], combustor and nozzles [43–45], contra-rotating turbines [39],
of scramjet engine design tools [8–11]. and high temperature heat exchangers [39]. The testing of an integrated
From 1996 to 2004, the NASA Hyper-X Program was initiated to airbreathing propulsion unit has been conducted at Mach 8 flight con­
demonstrate and validate critical scramjet engine technologies [12–21], ditions [46,47], where the feasibility of a positive aero-propulsive scaled
experimental techniques [22–28], and the computational methods and model in flight was verified.
tools [29]. This program was dedicated to designing and developing an The HyShot Flight Program, a rocket-boosted flight experiment for
advanced, airframe-integrated, airbreathing, hypersonic propulsion supersonic combustion investigation, was led by the University of
system via several small, relatively low cost, scramjet powered, auton­ Queensland (UQ) in Australia during 2001–2007 [48–50]. The realistic
omous flight demonstrator vehicles at Mach numbers 7 and 10 based on flight database of supersonic combustion was developed to correct the
past vehicle studies from the NASP database [7,17,30–33]. Three ground testing data obtained from the T4 shock tunnel at UQ. Pressure
unpiloted hydrogen-fueled X-43A research aircraft (Fig. 1a) were built measurements were the primary means for obtaining the correlation.
during the Hyper-X program. The first attempt failed in June 2001 due There were four flight experiments conducted during the program and
to boost rocket malfunction [17,34]. The record-breaking high-speed the supersonic combustion was achieved successfully in the HyShot II
atmospheric flights were achieved at Mach 6.8 speed (March 2004) and and III flights. The HyShot II flight data formed a foundational bench­
Mach 9.6 speed (November 2004) with hypersonic flight operation for mark for both experimental and numerical investigations of supersonic
about 12 s. combustion [51–57]. Inspired by the success of the HyShot Flight Pro­
From 2003 to 2013, the U.S. Air Force Research Laboratory (AFRL) gram, a collaborative effort between the United States’ Defense
and the Boeing Company carried out a series of flight tests on the X-51 Advanced Research Projects Agency (DARPA) and Australia’s Defense
Waverider (Fig. 1b), an unmanned scramjet aircraft for hypersonic flight Science and Technology Organization (DSTO) was established by the
at Mach 5. This experimental vehicle was used as a research platform for HyShot team, i.e., the Hypersonic Collaborative Australian/United
practical applications of hypersonic flight, such as a missile, recon­ States Experiment (HyCAUSE) program [50]. The HyCAUSE program
naissance, transport, and airbreathing first stage booster for a space tested an inward-turning scramjet engine concept in flight at Mach 10
system [35,36]. The main objective of the X-51A program was to (i) conditions. The ground testing was conducted in the T4 facility at UQ
acquire ground and flight data on an actively cooled, self-controlled and the LENS I (Large Energy National Shock Tunnel I) facility at CUBRC
operating scramjet engine for better understanding of the governing (Calspan-University of Buffalo Research Center).
physical phenomena and computational design tools used for scramjet The Hypersonic International Flight Research Experimentation
design, (ii) demonstrate viability of an endothermically fueled scramjet (HIFiRE) program was a collaborative international effort between the
in flight, and (iii) produce greater thrust than drag to prove the viability U.S. AFRL and Australia’s DSTO [58–60]. The HIFiRE program aimed to
of a free-flying, scramjet-powered vehicle using endothermic hydro­ investigate fundamental hypersonic phenomena at Mach 6.0–8.5+
carbon fuel at Mach number 4.5–6.5. In May 2013, the X-51 hypersonic through realistic flight experiments, enabling research in flight regimes
vehicle finally achieved the peak speed of Mach 5.1 with a record-setting that are expensive and difficult to model/test with existing codes and
210 s of airbreathing hypersonic flight on the fourth flight, the longest ground test facilities [60,61]. Up to ten flights were planned to study
hypersonic flight under its scramjet power. Three experimental flight basic hypersonic phenomena in realistic flight regimes using sounding
tests failed during 2010–2012 due to the inlet unstart, failure of fuel rockets to boost the experiments to the flight conditions of interest [62]
transition operation, and/or loss of aerodynamic control [37]. since 2007. The HIFiRE program followed the HyShot and HyCAUSE
In Europe, technology driven programs, under Long-Term Advanced programs and leveraged many of the low-cost flight test techniques
Propulsion Concepts and Technologies (e.g., LAPCAT-I and LAPCAT-II), developed in those programs [61]. The program conducted a series of
have been carried out since 2005 with the mission of reducing travel flight tests HIFiRE 0–7 encompassing a wide spectrum of
times of long-distance flights, which required investigating new flight disciplinary-specific research problems, ranging from the determination

Fig. 1. Hypersonic airbreathing vehicles (a) X-43A (U.S. NASA), (b) X-51A (U.S. Air Force).

2
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

of boundary layer transition onset (HIFiRE-1, HIFiRE-5) [63], evalua­ 91]. These ground test facilities generally achieve either
tion of hydrocarbon fuel combustion kinetics (HIFiRE-2) [60], inlet high-temperature and high-pressure clean flow conditions for a short
starting (HIFiRE-2) [64], radical farming (HIFiRE-3) [63], thrust mea­ duration, or contaminated, limited pressure and enthalpy conditions for
surement (HIFiRE-7) [65], and guidance and control (HIFiRE-0, a long test duration. There is always an implied compromise when
HIFiRE-4, HIFiRE-6) [63,66]. The HIFiRE-8 [67,68] was planned as a testing in the hypersonic regime. It is noteworthy that Chinese Academy
scramjet-powered waverider vehicle to fly horizontally for up to a of Science has developed the Long-test-duration Hypervelocity
minute at Mach 8 but the program ended prior to its launch. Detonation-driven Shock Tunnel (LHDst) with test time (~100 ms)
These representative research projects on hypersonic airbreathing much longer than the conventional shock tunnel facilities to bridge the
propulsion played a key role in advancing national security, such as the gap between the demanded flow parameters and characteristic time [92,
Prompt Global Strike (PGS) hypersonic weaponry [69]. The PGS is a 93].
strategic military effort implemented by the US to deliver a Despite the fact that the hypersonic experiments based on ground
precision-guided conventional weapon airstrike anywhere in the world test facilities have been successful in providing fundamental under­
within 1 h, which allows swift and immediate response to rapidly standing and information essential for successful scramjet propulsion
emerging threats than is possible with conventional forces. Many op­ system designs and realistic flight tests (e.g., X-43 and X-51), laboratory
tions are considered in the PGS system, which includes the hypersonic testing is still a challenging task given the requirements of duplicating
cruise missile (HCM) and hypersonic glide vehicles (HGV). HCMs fly extreme pressures, temperatures, and a range of nondimensional simi­
faster than intercontinental ballistic missiles (ICBMs) and have the larities that are beyond the physical limitations of ground test capabil­
maneuverability of a cruise missile. HGVs are ground launched by a ities [71]. For instance, the mode transition operation of scramjet
rocket, deploy at high-altitudes, and finally glide to strike targets. The engines during the accelerating phase is difficult to study in
hypersonic airbreathing vehicles provided potentially the most efficient ground-based facilities since most facilities operate at fixed Mach
and the longest range delivery system for PGS warheads, such as the air- numbers and the transition process involves highly unstable reactive
or submarine-launched hypersonic cruise missile and unmanned flows, the nature of which can be obscured by chemical composition of
waverider. the test gas that differs from actual flight [62]. Therefore, actual flight
The design and development of hypersonic vehicles over the past 60 tests such as the HIFiRE flight programs [63] are critical in validating
years has been an evolutionary process, in which progress was based on the science acquired in ground test facilities.
previous experience and databases accumulated through ground test
facilities and flight tests [70]. To facilitate the understanding of funda­
1.2. Supersonic combustion stabilization
mental physics of airframe/aerodynamic and propulsive integration
required for successful hypersonic airbreathing designs, various ground
High-speed airbreathing vehicles require stable combustion re­
testing facilities have been developed to produce hypersonic flows with
actions within the engine to achieve reliable thrust and high propulsion
designed flow enthalpy, Mach number, dynamic pressure, flow quality
efficiency over a wide range of flight Mach numbers (M∞ = 5–25) and
(turbulence level and contamination), and sufficient test time to simu­
altitudes (20–55 km) [1]. In hypersonic flight, the flow residence time in
late unsteady flow and combustion dynamics as well as thermal material
the scramjet combustor is approximately 10− 4–10− 3 s, whereas the
behaviors. Because of the combined affordability and physical limita­
chemical reaction time scale, which depends on flow temperature,
tions, ground test facilities can only partially replicate the correct flow
pressure, and mixture composition, varies from 10− 10 s to more than 1 s
conditions (e.g., temperature, pressure, velocity and chemical compo­
[94]. Furthermore, fast chemical processes (equilibrium conditions) and
sition) within the target flight regimes for research, test, evaluation, and
long chemical processes (frozen conditions) seldom exist in the scramjet
design tool validation purposes [71].
engine. The comparable time scales of flow and combustion lead to
The key capabilities of ground test facilities in most recent hyper­
closely coupled fluid/combustion dynamics and unfavorable conditions
sonic experimental requirements involve high Mach numbers, high en­
for combustion stabilization. When instability is increased, the flames
thalpies, high pressures, and longer run times. These research-oriented
can blowout or propagate upstream until inlet unstart occurs, which
facilities for fluid and combustion dynamics studies include but are not
leads to sudden loss of thrust and/or in-flight engine malfunction [37,
limited to 1) the hypersonic quiet wind tunnels for hypersonic boundary
95]. Hence, hypersonic vehicles demand effective strategies for com­
layer transition, such as the Boeing/AFOSR Mach-6 Quiet Tunnel
bustion stabilization and flameholding in airbreathing high-speed
(BAM6QT) at Purdue University [72,73], the NASA Mach 6 Quiet
environments.
Tunnel at Texas A&M University [74], and the Large Mach 6 Quiet
In such conditions, combustion needs to be anchored inside the
Tunnel at University of Notre Dame [75]; 2) the shock tunnels and
combustor to complete the heat-release and recombination reactions
expansion tubes providing high Mach numbers, high flow enthalpies,
after encountering a series of rate-limiting processes along the flowpath
high dynamic pressures, and dynamically near-clean air within limited
including fuel jet injection, fuel-air mixing at molecular levels, fuel
test times (order of a millisecond) for fundamental aerodynamic, aero­
ignition, flame propagation, and combustion stabilization within the
thermodynamic and combustion research, such as the expansion tube
combustor length. The criterion for scramjet combustion stabilization
facilities at Stanford University [76], California Institute of Technology
can be represented by:
[77], and University of Michigan [78], the High Enthalpy shock tunnel
/
Gottingen (HEG) facility at the German Aerospace Center (DLR) [79], (τmix + τchem ) τflow ≤ 1, (1)
the T4 free piston shock tunnel at University of Queensland [80,81], and
the X3R reflected shock tube at the Defense Science and Technology where τmix, τchem, and τflow are characteristic time scales for supersonic
(DST) group in Australia [82]; and 3) the long duration test facilities flow mixing, combustion chemical reactions, and residence flow time in
with sufficient run time (order of hundreds of milliseconds to contin­ the combustor chamber, respectively. Flame stabilization in high-speed
uous) for studies of basic mechanisms governing the mixing and com­ flows needs to provide means to (i) decrease the flow mixing time τmix,
bustion process and combustion-driven fluid dynamic phenomena, such (ii) elongate the flow residence time τflow, and (iii) reduce the combus­
as, the combustion-heated supersonic research facilities (Research Cells tion chemical reaction time τchem.
18, 19, 22) at U.S. Wright-Patterson Air Force Base (WPAFB) [83–87], At the low corridor of hypersonic flight, i.e., flight Mach numbers 5 <
the electric ohmic-heated supersonic blowdown rig (SBR-50) at Uni­ M∞ < 8, combustion stabilization may become a predominant issue due
versity of Notre Dame [88], the University of Virginia Supersonic to the long fuel ignition delay associated with the low stagnation tem­
Combustion Facility (UVASCF) [89], and the pulsed arc-heated com­ perature at the combustor entrance that may not be high enough to
bustion tunnel (ACT-II) at University of Illinois, Urbana-Champaign [90, ensure ignition within a reasonable distance downstream of the fuel

3
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

injection ports [96,97]. The conventional approach of scramjet com­ other scramjets including milder compression ratios to alleviate tem­
bustion is to mix and burn the fuel simultaneously, and this diffusive perature rise at combustor entrance, enhanced fuel/mainstream air
combustion mode has been the focus of the present-day high-speed mixing at the combustor entrance, reduced combustor lengths and hence
propulsion research. less skin friction drag, and reduced fuel ignition times. In this configu­
For diffusion-dominated combustion in scramjets, flameholding ration, the fuel is injected into the inlet to achieve a partially premixed
schemes have been proposed typically based on passive and active fuel-air mixture at the combustor entrance. By moving the injection
control of the characteristic fuel-air mixing time and flow residence upstream, fuel and air are provided additional length in which they can
time, which are the rate-limiting physical parameters during the com­ mix and burn within the engine. Along the flowpath of fuel/air mixtures
bustion process. The characteristic time of fuel-air mixing can be flowing from the inlet, the radicals required for ignition and combustion
reduced by introducing large-scale flow structures for passive mixing are formed in hot pockets that are high pressure and temperature re­
enhancement, e.g., shock-enhanced mixing, three-dimensional (3D) jets, gions between shock/expansion flow structures in the combustor and
axial vorticity generation, and excitation of shear layer instability. are coupled with the combustion reactions. The radicals are convected
Large-scale coherent structures at fuel-air interfaces enhance the mixing downstream and ignition is achieved in stages via two or more of such
process through entraining massive amounts of oxidizer into fuel rich hot pocket regions [118]. This mechanism leads to the shortening of the
regions, stretching fuel-air interfaces, increasing the interfacial areas, time required for the combustion chemical reactions, potentially
and/or steepening the local fuel concentration gradients in the molec­ reducing the fuel-air mixing time and increasing the fuel residence time
ular diffusion process [86,98–100]. Meanwhile, the efficient fuel-air compared to non-radical farming scramjets of the same configuration.
mixing in diffusion flame promotes regions of rich and lean premixed The flame is stabilized at supersonic speeds by the conveniently located
conditions that are favorable for flame propagation upstream to balance oblique shocks. The combustion may occur far downstream without
the approaching freestream speeds [101]. The active methods for mixing perturbing the preceding shockwave, i.e., shock-induced combustion;
enhancement include pulsed jet injection [102], pulsed detonation otherwise, the combustion process can couple with shockwaves with
[103], and plasma-intensified mixing techniques [104,105], which are extremely fast reactions forming oblique detonation waves (ODWs)
controlled by external energy deposition into combustion systems. On [119].
the other hand, recesses in the walls of scramjet combustors, i.e., steps At higher flight Mach numbers of hypersonic vehicles (M∞ > 12), the
and cavities, have been used as flameholding devices since they can concept of shock-induced combustion ramjet engine (shcramjet), or the
stabilize and pilot the core combustion by creating a low-speed recir­ ODW engine has been proposed for hypersonic airbreathing propulsion
culation region with favorable flame stabilization conditions in the su­ applications. To stabilize combustion at these extremely high speeds, the
personic flow, leading to an increase in the overall flow residence time performance of conventional scramjet designs is enhanced by replacing
[106]. The flameholding may be achieved by partial subsonic combus­ the long scramjet combustor by a much shorter one that employs a very
tion within the recirculation flow because a stable flame existing in the thin standing oblique detonation wave attached to a wedge [120–122].
cavity may act as an ignition source for the core flow and provide a As such, the fuel/air mixing and combustion process are decoupled. Like
continuous hot radical supply to reduce the fuel ignition induction time, the radical-farming scramjet engines, the fuel/air mixing is organized in
and hence act to enhance/stabilize the combustion downstream [100]. the vehicle’s forebody, in which the engine inlet may compress free­
Step and cavity flameholders are attractive means of combustion sta­ streams less in order to avoid or delay the ignition of the fuel/air
bilization in scramjet engines at low flight Mach numbers due to their mixture. The combustion of the approximately homogeneous gas
minimal impact on the supersonic core flow, including the flowpath mixture is achieved by an oblique shock that rises the temperature and
total pressure losses, reduced drag, and aerodynamic heating, despite pressure to its ignition point. The potential advantages of the
the large local wall heat fluxes [97]. External energy addition using detonation-wave ramjet engine over the scramjet engine include the
plasma and electric discharge [88,104,105,107–115] can kinetically reduction of total pressure losses during the inlet flow deceleration and
enhance and stabilize flames in high speed flows via plasma-induced compression processes due to the additional strong compression by the
ignition, plasma-intensified mixing, and flameholding by plasma gen­ detonation wave, shorter combustor length due to the rapid
eration [116], i.e., reducing the mixing and combustion chemical reac­ shock-induced combustion, less combustor cooling load, and an overall
tion characteristic time scales. shorter and lighter-weight engine system [123].
At the moderate corridor of hypersonic flights, i.e., flight Mach A regime diagram of scramjet combustion stabilization techniques is
numbers 8 < M∞ < 12, the flameholding methods proven to be effective proposed in Fig. 2. It classifies the most common categories of com­
in low-to-moderate hypersonic freestream Mach numbers may cause bustion stabilization methods based on their contribution to mixing and
significant total pressure losses in the flowpath, resulting in engine chemistry enhancement. It is understood that in order to achieve effi­
performance degradation with increasing flight Mach number. Mixing cient combustion, it is necessary to (i) achieve sufficient mixing of fuel
deterioration at high Mach number flows due to the compressibility and oxidizer at the molecular level, and (ii) initiate the chemical chain
effect and the limited active zone of external energy addition prevents reaction sequence overcoming the activation energy. The first require­
the volumetric consumption of the compressed freestream, particularly ment is usually fulfilled by producing turbulence in the flow that in­
in the supersonic core. The overall performance of scramjets may be creases the contact surface between fuel and oxidizer streams,
reduced drastically by the overall length required to complete the accelerating the molecular diffusion process. The second requirement
combustion, vehicle frictional drag, and heat transfer [117]. In addition, can be accomplished by two distinct but synergetic mechanisms: one
due to the increasing flight speed, the elevated temperature at the thermal, consisting of a rise in the local flow temperature, and one
combustor entrance will dissociate the combustion products resulting in chemical, consisting of the production of radicals and excited species.
reduced total energy release and water content from the combustion at Both of these mechanisms act to accelerate the chain reaction process
the exit of the combustor, leading to significantly reduced specific im­ but in different pathways. Increasing the temperature directly increases
pulse [117]. A possible combustion stabilization and performance the reaction rates, according to Arrhenius equation, whereas seeding the
enhancement method of the scramjet engine is to manipulate the flow with highly reactive particles allows bypassing the higher activa­
chemical reaction time by converting diffusive combustion to partially tion energies associated with chain starting reactions, ultimately
premixed combustion occurring in engine combustors within a reason­ resulting in decreased ignition delay time.
able overall length. Every practical combustion enhancement/stabilization method
The inlet-fueled radical-farming class of scramjets was proposed and contributes to some extent to both these aspects. Nevertheless, each
experimentally tested at the University of Queensland, Australia [80]. category is more effective in pursuing either chemistry or mixing
These scramjets have been shown to offer many merits compared to enhancement, and for this reason, each one can be optimized for

4
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 2. Regime diagram of combustion stabilization methods for hypersonic propulsion.

different applications and different flow regimes. Vortex generators, chemical reactions [128]. Experiments [129] show that the piloting
struts, etc. produce a highly turbulent wake that promotes mixing with flame using hydrogen can extend the flameholding ability of a
minimal contribution to local temperature rise. Therefore, they can be kerosene-fueled cavity. The trend of combustion stabilization method as
considered as pure mixing devices, which are more effective in the case shown in Fig. 2, i.e., transitioning the diffusive combustion to partially
of mixing-limited reactive flows. Steps and cavities provide large regions or fully premixed combustion at high Mach number conditions, would
of recirculating flow and shear-layers where mixing is enhanced and not be changed by the gaseous/liquid fuel properties.
regions of low-speed flow where the recovery temperature is sufficiently
high to impact the chemistry. Plasmas and other energy deposition de­ 1.3. Scope of the review
vices act mainly through thermo-chemical effects. In a discharge for
example, the local temperature of the gas is increased by Joule heating Despite the considerable progress made in the development of the
when electric power is dissipated into the plasma column, and ions and scramjet propulsion technology over the past 60 years, the under­
other long-lived radical species are seeded into the flow and transported standing of the physics involved in high-speed reacting flows is still far
into reactive zones. It has been shown [88,104,105,115] that plasma from complete [2]. Recently, many review articles have been published
assisted techniques also introduce a significant mixing component. The on supersonic combustion in airbreathing engines for hypersonic flights.
large gradients created by the plasma discharge act as sources of tur­ Urzay [130] surveyed recent experimental flights and ground-based
bulent kinetic energy and vorticity by baroclinic effects and can be research programs, addressing the fundamental flow physics and nu­
exploited to accelerate flow mixing. On the other hand, the predominant merical modeling of supersonic combustion. Wang et al. [96], Barns and
effect of shock induced combustion is thermal. The idea is to ignite an Segal [97] emphasized cavity-based combustion stabilization, flame
already sufficiently mixed flow by suddenly increasing its temperature stability analyses, and combustion oscillations, providing comprehen­
beyond the auto-ignition limit. In this case, it is crucial that the mixture sive knowledge of cavity flameholding mechanisms in the low hyper­
composition is already within the flame ignition limits because the sonic flight regime. Chang [131] summarized research on strut-based
mixing effect introduced by a shock is very limited and confined to a supersonic combustion. Although these reviews have provided excellent
region close to the solid boundaries, where the shock interacts with the overviews of distinctive processes of hypersonic propulsion, the under­
local boundary layer. For this reason, shock induced combustion in lying general logic and science of combustion stabilization over the
scramjets is usually accompanied by an early injection of the fuel, which entire hypersonic flight envelope from low to high Mach numbers has
increases the overall flow residence time, or by mixing enhancing yet to be outlined. The goals of this review are to: 1) provide a general
devices. logic of underlying physics on supersonic combustion stabilization, 2)
The methods used to achieve the combustion stability under different present an overview of combustion stabilization techniques in a wide
Mach numbers are categorized qualitatively based on their abilities to range of high Mach number flight regime, 3) provide a comprehensive
enhance mixing and chemistry as well as total pressure losses with the understanding of important research progress on diffusive and
increasing Mach number. It is noteworthy that different fuels and their shock-induced combustion modes, 4) discuss design guidelines in a
proportions have impacts on flame stability. For instance, liquid fuel scramjet combustor to achieve combustion stabilization, and finally 5)
sprays have droplet breakup and vaporization processes before mixing discuss the challenges and needs of future research in hypersonic com­
with approaching freestreams [124–127]. Regenerative cooling can bustion stabilization.
crack the hydrocarbon fuel into small molecules to promote fast

5
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

2. Diffusive combustion simplified conditions with dominant controlling mechanisms for flame
structure. Reviews of these idealized supersonic combustion models
Diffusive combustion has been widely studied for high-speed pro­ were provided by Ferri [132] and Curran et al. [133]. Recent research on
pulsion systems at the lower corridor of hypersonic flight. In this type of supersonic combustion has been focused on resolving the micro-scale
combustion configuration, fuels are injected into scramjet combustors mechanisms related to detailed flame structures and heat release rate
for simultaneous diffusive mixing and burning, which results in using both experimental and numerical methods. Summaries of the main
partially-premixed or diffusive flame modes. The supersonic combustion physical aspects and the progress on numerical simulations of full
in scramjet combustors is achieved through an extremely complex scramjet systems are provided by reviews from Gonzalez-Juez et al.
aerodynamic process where the chemical and fluid dynamic effects [134] and Urzay [130]. The practical problems occurring in scramjet
strongly interact at high speeds on the order of 1000–3000 m/s [130, combustors involve complex physical processes including transport
132]. The early studies on scramjet combustion were focused on the mechanisms and chemical kinetics, compressible turbulence and mix­
fundamental combustion modeling in compressible reacting flows for ing, flow recirculation, shock-enhanced mixing and combustion,

Fig. 3. Flow and flame parameters in an ethylene-fueled scramjet combustor [101].

6
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

thermal choking, and near-wall burning [130]. The totality of these field mixing efficiency with moderate total pressure losses and me­
problems results in significant challenges in analyzing the flame struc­ chanical complexity for practical engineering applications. In general,
ture and predicting the flame anchoring location. fuels are delivered into the supersonic freestream at sonic or supersonic
In scramjet engines where fuel is injected close to or directly into the speeds using simple jet injection methods, which include transverse
combustor, there exist partially-premixed or fully-premixed flames injection to crossflow, parallel injection to crossflow, and oblique in­
locally, though the general combustion occurs in a diffusive mode. The jection into crossflow. Fig. 4 shows the schematics of jet injection
portion of the partially premixed flame has been the rate-limiting schemes. It is noted that the fuel injection angle can be normal to the
parameter for overall combustion efficiency. Potturi and Edwards airstream at M∞ < 10 to enhance penetration but must approach coaxial
[101] numerically investigated the ethylene flame dynamics in a Mach 2 at high M∞. At flight speed M∞ > 10, the momentum of the fuel becomes
cavity-based supersonic combustor. Fig. 3 shows the generalized flame an increasingly important element in the thrust potential of the engine
index contours that indicate the distributions of diffusion flames along [117].
with lean- and rich-premixed flames. The flame initially propagates as a
premixed flame into a slightly rich fuel-air mixture inside the cavity, 2.1.1. Transverse jet injection
after which the flame propagation is inhibited near the stoichiometric Transverse fuel jet injection into supersonic cross flows has been
line. Lean-premixed and diffusion flames, which are characteristics of considered a simple and effective fuel delivery scheme in supersonic
partially premixed flames, start to dictate the combustion regions combustion engines because the jet-crossflow interactions introduce
downstream of the cavity trailing edge. The maximum heat release oc­ efficient fuel and air mixing in the near field of the fuel jet [100,146].
curs at the transition from a rich premixed to a partially premixed flame. Single injector and/or arrays of transverse injectors have been used as
Hence, the flame stabilization of the diffusive combustion in scramjet fueling schemes for ramjet and scramjet engines. Interactions between
combustors has been the combined response of the combustion process the underexpanded transverse jet and supersonic crossflows have been
to the harsh flow conditions. Nevertheless, it is challenging to experi­ extensively investigated in reacting high-enthalpy flows to understand
mentally investigate the associated flow structures, fuel-air mixing, fuel the mechanisms of fuel-air mixing and flame stabilization.
ignition, and combustion stabilization characteristics in high-speed and Fig. 5 shows the instantaneous flow structures and the averaged 3D
high-temperature supersonic freestreams in realistic flight flowfield of transverse jet injection into supersonic crossflows. The
environments. underexpanded jet forms a complex system of expansion waves and
The key to stabilizing diffusive combustion in scramjet combustors is shockwaves (e.g., barrel shocks and Mach disk) due to deformation by
to promote the proportion of partially premixed flames. Physical pro­ hydrodynamic forces as the jet penetrates the supersonic crossflow. The
cesses involved in the promotion of partially premixed flames in the near fields of transverse jets in supersonic crossflow are characterized by
diffusive combustion mode include [99]: (1) formation of coherent 3D bow shocks generated by supersonic crossflow blockage, flow
structures containing unmixed fuel and air, wherein a diffusion flame recirculation regions confined by a separation shock at the windward
occurs as the gases are convected downstream, (2) interaction of a shock side of the jet, jet-shear layer vortices in the windward side of the
with partially mixed fuel and oxidizer, (3) organization of a recircula­ transverse jet, and a horseshoe vortex region wrapping around the jet
tion area where the fuel and air can be mixed partially at low speeds, and
(4) chemical kinetic modification by external energy addition. In this
section, the details of major research efforts for diffusive combustion
stabilization will be reviewed, and the challenges will be discussed for
successful combustion stabilization in scramjet engines for high speed
propulsion.
Flow-path design of the scramjet engine is very crucial for the
improvement of its combustion stability and overall combustor perfor­
mance. Some representative numerical data such as secondary flow path
or flow path with cavity are supplemented here including HIFiRE-2
flowpath [135–138], HyShot II flowpath [56,57,139], JAXA (Japan
Aerospace Exploration Agency) M6 engine flowpath [140], scramjet
flowpath from the University of Virginia [89,141], and REST (Rec­
tangular-to-Elliptical Shape Transitioning) flowpath from the University
of Queensland, Australia [142–145].
For hypersonic airbreathing propulsion, the flight Mach number and
airstream Mach number at the entry to combustor are different. The
freestream speed is decelerated by the inlet compression system and
shock trains in the isolator. For instance, for practical scramjet engines,
the Mach 7 flight speed leads to around Mach 3 flow speed at the
combustor inlet, whereas Mach 5 flight corresponds to Mach 2 flow
speed at the combustor inlet. For experimental investigation focused on
combustor dynamics, a direct-connect wind tunnel configuration is
usually used. The Mach number of the wind tunnel freestream needs to
match the combustor inlet conditions, which is under Mach 5. Therefore,
most of the data obtained from diffusive-type supersonic combustion
investigation are under Mach number 5 freestream condition.

2.1. Simple fueling schemes

In the diffusive combustion configuration, a scramjet combustor re­


quires efficient molecular level mixing of fuel (hydrogen or hydrocarbon
fuels) and air in the limited flow residence time and combustor length at Fig. 4. Scramjet fueling schemes: (a) transverse, (b) inclined, and (c) parallel
hypersonic flight speeds. Fueling schemes need to maximize the near- injections. Red indicates mixing and reaction, blue indicates fuel rich region.

7
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 5. An underexpanded transverse injection into a supersonic crossflow: (a) 3D view of density gradient [192], schematics of (b) the instantaneous jet-crossflow
interaction [100], and (c) the averaged flowfield [98].

envelope. In the far field, the flowfield is characterized by a streamwise general structure and overall performance prediction of the fueling
counter-rotating vortex pair formed by the transformation of the jet- system, due to the limited capability to visualize the flowfield, such as
induced shear layer into a folded vortex sheet and wake vortices low-speed schlieren photography [150,156–164]. Advancements in
formed by the horseshoe vortex. The flow recirculation regions at the laser diagnostic techniques has enabled access to the detailed flowfield
windward side of the jet enable the mixing between the boundary layer mixing characteristics and flame structures of jet-crossflow interactions
and jet fluids and may act as a continuous fuel ignition source in high- using non-intrusive optical measurements both qualitatively and quan­
enthalpy flows. The mixing properties of the transverse jet in super­ titatively, such as laser Doppler velocimetry (LDV) [165], particle im­
sonic flows are controlled by the jet vortical structures that are corre­ aging velocimetry (PIV) [166], Mie scattering [98], planar laser
lated with the jet-to-freestream momentum flux ratio and the evolution Rayleigh scattering (PLRS) [167], nanoparticle-based planar laser scat­
of the jet shear layer under large velocity gradients [98,100]. In general, tering technique (NPLS) [149], Raman scattering [168,169], and planar
streamwise vortices induce large-scale flow structures to entrain massive laser-induced fluorescence (PLIF) (iodine [170], OH [146,155,
amount of oxidizer into the fuel, stretch fuel-air interfaces, increase the 171–178], nitric oxide [171,172,179], acetone [180–182], toluene
interfacial areas, and/or steepen the local concentration gradients to [152]). These laser diagnostics techniques are also used to characterize
enhance the molecular diffusion process, i.e., the tilting-stretch-tearing the flow and combustion dynamics of other flame stabilization methods
mechanism of fuel-air mixing in the transverse jet fueling configura­ in high Mach number flows. To fully resolve the turbulent flame struc­
tion. Summaries of the jet-crossflow interaction in both subsonic and ture and flow dynamics in supersonic flows, high-speed imaging systems
supersonic flow regimes can be found in Refs. [147,148]. need to reach a frequency of 100 kHz and above. Recent advances in
Extensive studies on transverse jets in supersonic crossflows have ultra-high repetition rate laser diagnostics system based on burst mode
been performed to determine the controlling parameters of the mixing lasers provide powerful tools for fluid and combustion dynamic mea­
mechanism, fuel ignition, and combustion characteristics, which include surements in hypersonic propulsion systems [183,184]. Additionally,
dynamics of large-scale coherent structures at the jet-shear layer inter­ many valuable studies using computational fluid dynamics (CFD)
face [98,100,149], jet-to-freestream momentum flux ratio [100,146], methods such as large eddy simulation (LES) and direct numerical
fuel molecular weight [99,100,150–152], boundary layer thickness simulation (DNS) methods provide further understanding of
[153,154], stagnation temperature [151,155], and compressibility ef­ jet-crossflow interactions in both non-reacting and reacting flow con­
fects [98]. Substantial experimental studies were conducted in figurations that cannot be visualized experimentally [185–191]. Table 1
non-realistic flow conditions (e.g., low flow enthalpy, pressure, velocity) lists selected studies on jet-crossflow interactions over the past two de­
to investigate various physical aspects, which heavily depend on the cades with an emphasis on laser diagnostics and LES methods.
capability of the ground test facility and the diagnostics tools deployed. Jet trajectory is considered the most significant feature of jet-
Early studies on transverse jets in supersonic flows were focused on the crossflow interaction to characterize the development of the large-

8
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Table 1
Experimental studies on transverse jet in supersonic crossflow.
Injectanta Transverse jet Freestream conditions Methods Topics Refs.

Ma Angle Ma P0 (bar) T0 (K)

Air 1.0 90◦ 1.6 2.41 295 LDV Vortex structure, shear layer [165]
Air 1.0 90◦ 5.0 6.5 375 PIV Velocity fields, jet mixing [166]
H2, CO2 1.0 15◦ ,90◦ 2.0 10.6 300 Mie scattering Coherent structure, mixing [98]
He, Air 1.0 90◦ 2.0 3.17 300 Mie scattering Compressibility effect [86]
CO2 3.4 60◦ 4.5 1.0 300 PLRS Jet structure and flowfield [167]
N2 1.0 90◦ 1.0 1.0 300 NPLS Vortex structure evolution [149]
C2H4 1.0 30◦ ,90◦ 2.0 2.4 300 Raman scattering Fuel plume concentration [168,
169]
N2/NO, H2 1.0 90◦ 1.4 1.3 2900 NO/OH-PLIF Mixing, flame distribution [171,
172]
H2 1.0 30◦ ,60◦ 3 11.2 4000 NO/OH-PLIF Mixing, flame distribution [174]
NO/CO/H2 1.0 90◦ 1.4 0.4 2200 NO/OH-PLIF Temperature [179]
H2 1.0 90◦ 1.5 1.0 3500 OH-PLIF Jet trajectory and penetration [173]
H2 1.0 90◦ 3.46 26.4 3790 OH-PLIF Jet flame structure [100]
H2 1.0 90◦ 2.26–2.8 2.90–7.15 2500–3100 OH-PLIF Total temperature effect on jet combustion stability [155]
C2H4 3.4 60◦ 4.5 1.0 2600 OH-PLIF Volumetric flame structure [194]
H2 1.0 90◦ 2.4 5.80 2800 OH-PLIF Fuel ignition, flame structure [146,
195]
Air 1.0 90◦ 1.6 2.41 295 Acetone PLIF Jet flow structure, mixing [180]
Air 1.0 90◦ 1.9 0.75 291 Acetone PLIF Concentration distribution [181,
182]
H2, He, Ar, N2 1.0 90◦ 2.3 13.9 900 Toluene PLIF Temperature and number density [152]
H2, C2H4 1.0 90◦ 3.38 23.6 3650 High-speed Dynamics of jet-shear layer coherent structures, jet trajectory [99,100]
schlieren
He, Air, C2H4 1.0 90◦ 2.0 2.4–3.0 300 LES Boundary-layer structure, mixing [185]
Air 1.0 90◦ 1.6 4.8 300 LES Turbulent boundary layer, unsteady jet-crossflow entrainment [186,
188]
Air 1.0 90◦ 1.6 2.4 295 LES Shock-turbulent boundary layer interaction, shear layer [187,
instability 196]
Air 1.0 90◦ 2.7 1.0 300 DNS Turbulence statistics of jet-to-cross-flow interaction [191]
H2 1.0 90◦ 2.48 6.0 3000 LES Jet autoignition in reacting flow [189]
a
All injectants are in room temperature, i.e., a total temperature of 300 K.

scale coherent structures at the jet-shear layer interface evolving along calculated and measured data set from assembled jet penetration
the flowpath. The jet penetration has been shown to be dependent pri­ studies. McClinton [153] investigated normal sonic injection of
marily on the jet-to-crossflow momentum flux ratio J, hydrogen into a Mach 4.05 airflow with the plate boundary-layer
( ) thickness to jet diameter ratio δ/d = 1.25–6.5. It was discovered that
(ρu2 )j γpM 2 j
J= = ( ) , (2) increasing δ/d increased penetration in the outer edge of the mixing
(ρu2 )∞ γpM 2 ∞ region and the mixing efficiency, which were respectively defined by the
0.5% jet concentration contour and the decay rate of the maximum
where ρ, u, p, γ, M are the density, velocity, pressure, specific heat ca­ hydrogen concentration. Portz and Segal [154] evaluated the penetra­
pacity ratio, and Mach number, respectively. The subscript j corresponds tion of transverse jets into supersonic crossflows at Mach 1.6–2.5 and
to the jet exit conditions and ∞ corresponds to the unperturbed free­ boundary layers δ/d = 0.8–3.7. The effect of the boundary layer thick­
stream conditions. The jet-to-crossflow momentum flux ratio J and the ness on the near-field jet penetration was significant at low Mach
jet orifice d are used as the normalizing factors in empirical non- numbers but the effect decreased with increasing Mach number. Gruber
dimensional correlations of jet trajectory from flow visualization. For et al. [86] explored sonic helium and air jet injection into a Mach 2
jets in the supersonic crossflow, various criteria have been used to supersonic crossflow at identical jet momentum flux ratios to examine
characterize the jet penetration, such as the height of the Mach disk from the effect of compressibility on jet penetration and mixing characteris­
schlieren imaging [150,151,156–159,161,162], the detectable upper tics. It was observed that the large-scale eddies formed in the shear layer
edge of the jet from schlieren imaging [99,100], the 0.5% boundary of dissipated rapidly in the air injection case and preserved coherence over
injectant concentration from translational gas sampling analysis [193], a longer spatial distance in the helium injection case. The difference of
the 1% threshold of fuel mole fraction from quantitative PLIF mea­ large-scale structures between air and helium jets were attributed to the
surement [170], the 90% intensity of scattering signal behind the bow compressibility effect, i.e., the convective Mach number Mc was three
shock from Rayleigh/Mie scattering diagnostics [86], and the 10% times large for helium compared to air. Portz and Segal [154] and
threshold of the maximum intensity of flame hydroxyl chem­ Gruber et al. [86] found that the influence of injectant molecular weight
iluminescence [146]. A compilation of correlations for jet trajectories in on jet transverse penetration into the crossflow was negligible.
reacting and non-reacting supersonic crossflows can be found in Refs. Ben-Yakar et al. [99,100] examined the reacting sonic hydrogen and
[146,148]. ethylene jets transversely injected into high-enthalpy supersonic flows
The boundary layer thickness δ, the convective Mach number Mc, at Mach number 3.38 and observed a deeper ethylene jet penetration
and the fuel molecular mass Mw have been characterized by their effects than the hydrogen jet with the same jet momentum flux ratio. This was a
on jet penetration as well as the jet-to-crossflow momentum flux ratio J. result of the tilting–stretching–tearing mechanism induced by the large
Povinelli [193] studied the transverse injection of a helium jet at Mach velocity gradient between the slow jet (the higher the jet molecular
1–4 into supersonic crossflows at Mach 2 and 3 to characterize the effect weight, the slower the jet exit velocity) and the fast free stream. i.e.,
of different parameters for a comprehensive correlation. It was found coherent structures tilted in the crossflow direction and lost their
that the addition of the boundary layer momentum thickness to the coherence more slowly at larger growth rates of jet shear layer (ethylene
correlation equation improved the agreement to ±15% between the injection). Additionally, jet penetration was found to have minimal

9
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

dependence on the freestream Mach number, jet Mach number, pressure increasing the interfacial surface area and the perimeter of the jet fluid.
ratio, and jet-crossflow density ratio [154,197]. The ensemble analysis of instantaneous images indicated characteristics
Time evolution and mixing characteristics have been experimentally of the lateral jet spread, the plume area, and the feature of fluctuation
investigated by resolving the detailed vortices and coherent structures intensity in the highly 3D flowfield, which are dominated by both the
existing in flowfields created by transverse jet injection into a supersonic large- and small-scale vortices in the counter-rotating vortex pair and jet
crossflow. Coherent structures at the jet-shear layer interface and mixing wake regions. The fuel-air mixing region wrapping around the jet en­
analysis have been studied using Mie or Rayleigh scattering imaging, velope increased along the flowpath, and the non-mixed jet core mixing
NPLS, and ultra-fast-framing schlieren imaging. Mie or Rayleigh scat­ was more uniform further downstream.
tering methods depend on the collection of light scattered by particles VanLerberghe et al. [180] studied the near-field mixing character­
and/or molecules suspended in the flowfield of interest [86]. NPLS vi­ istics of an underexpanded sonic air jet injected into a Mach 1.6 cross­
sualizes the flowfield based on Rayleigh scattering and solves the fidelity flow. The instantaneous PLIF images from acetone molecules were used
issue of tracer nanoparticles particles in supersonic flows. As the nano­ to extract image-intensity probability density functions (PDFs) to
particle seeding is independent from the local thermodynamic condi­ quantify mixing and to develop a qualitative description of the
tions, potential quantitative mixing interpretations can be made from entrainment process. The large-scale coherent rolling structures that
the NPLS images. Higher signal intensity could be produced by the NPLS were observed to periodically pass through the mixing layer shown in
method than could be produced by the traditional molecule-based Fig. 7 were consistent with the trends shown by the PDFs, i.e., the
Rayleigh scattering methods [149]. The ultra-fast-framing schlieren rolled-up layers of unmixed jet fluid and unmixed crossflow fluid
imaging at megahertz repetition rate can temporarily resolve the evo­ transported by the large coherent structures yielded the high-probability
lution of jet-vortex structures along the edge of the transverse jet in wings of the PDFs. The PDFs indicate that significant instantaneous
high-speed flows [100]. near-field mixing occurred in the wake region downstream of the barrel
Gruber et al. [86] investigated the large-scale structural develop­ shock region and below the jet centerline, whereas large regions of
ment and the near field mixing characteristics in Mach 2 flow in the unmixed fluid is shown elsewhere. The significant difference between
RC-19 facility using planar Rayleigh/Mie scattering by silicon dioxide the rate of mixing at the top and bottom of the jet were attributed to the
particles seeded in the supersonic freestream. The instantaneous images velocity differences, convective Mach numbers associated with the top
provide information regarding flowfield structure evolution and are of the annular shear between the jet and crossflow, and vertical motion
shown in Fig. 6. The large-scale eddies in the interfacial region between associated with the streamwise/counter-rotating vortices. Quantitative
jet and freestream mainly developed at the upper jet envelope, while the LDV measurement carried out by Santiago and Dutton [165] confirmed
predominantly smaller structures developed at its sides. The large that the top shear-layer regions had relatively large Reynolds shear
extrusion feature from the jet fluid into the freestream may potentially stresses, whereas the bottom leeward side of the barrel shock had small
promote the diffusion of freestream fluid across the strained interface by Reynolds stresses. The study also showed that the fluid immediately

Fig. 6. Instantaneous end view planar Mie scattering images of (a) air and (b) helium jets in a Mach 2 airflow [86].

10
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 7. Instantaneous images of the side-view barrel shock region at the middle plane [180].

downstream of the leeward side of the barrel shock accelerated rapidly translate with little rolling, i.e., losing its ability to engulf the freestream
in an area of low transverse velocity gradient, which suggests that a for mixing.
stream of unmixed crossflow fluid wraps around the circumference of Ben-Yakar et al. [99,100] visualized the consecutive entrainment of
the inner jet core and impinges on itself in the region between the wall crossflow fluid by large-scale coherent structures and the transport of
and the jet plume. This is an important mechanism for mixing due to the flow coherent characteristics. The experiment compared the coherent
fluid entrainment caused by the counter-rotating vortex pair. motion of sonic hydrogen and ethylene jets in a Mach 3.38 airflow in an
Sun et al. [149] analyzed the evolution of the coherent structures at expansion tube facility. The evolution of coherent structures at the edge
the jet-crossflow interface using NPLS in a direct-connect supersonic test of the jet-shear layer interface was resolved by megahertz-framing-rate
facility at Mach 2.7 airflow with a non-reacting configuration. As shown schlieren imaging as shown in Fig. 9. Large-scale coherent structures
in Fig. 8, small-scale eddies were both directly transported from the from jet-shear layer vortices under large velocity gradients were part of
turbulent boundary layer upstream and produced at the jet-shear layer the unsteady Kelvin-Helmholtz circumferential rollers, whose cores
interface in the jet-bending region. At the near field of jet injection, the coiled up around the jet with their legs connected downstream of the jet
convection velocities were low due to the exchange of mass, momentum, exit. The eddy formation frequency scaled linearly with the exit velocity
and energy between the jet and crossflow fluids. The vortex growth rates of different sonic jets. Large fluctuations of the convective velocity in the
of the shedding eddies were high, and the small structures accelerated shock-induced freestream behind the rugged bow shock impacted the
rapidly along the flowpath. In the downstream region, the large-scale tilting and stretching of eddies in the mixing process. Convection ve­
eddies had less deformation and translated at velocities near the free­ locities and inclination angles were extracted from the visual image to
stream velocity. The jet eddies were mainly formed in the near field, and characterize the transportation of coherent structures. Analysis showed
the dimension of eddies in the near field correlated with the size of that the eddies persisted a certain distance while moving downstream,
eddies in the far field. It is concluded that the mixing process in the near and the dissipation rate varied with the shear stress due to the velocity
field determines the overall mixing rate since the eddies in the far field gradient. In high-speed freestreams, these large-scale coherent

Fig. 8. Instantaneous NPLS images of nitrogen jet penetration into a supersonic crossflow (J = 3.5) at the jet central plane with a frame gap time of 10 μs [149].

11
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 9. Consecutive schlieren images of underexpanded hydrogen injection (d = 2 mm) into a supersonic crossflow (nitrogen) obtained by high-speed-framing
camera [99,100]. Exposure time of each image is 100 ns and interframing time is 1 μs.

structures, in which the fuel and air are mixed by slow molecular 1.6 airflow at a stagnation temperature of 300 K. As shown in Fig. 10, the
diffusion, will also travel at high speeds. Therefore, the overall diffusive unsteady flowfield illustrated that the pressure fluctuation inside the
combustion process will be mixing controlled [100,132]. recirculation region upstream of the jet was coupled with the large-scale
Using experimental data provided by the fundamental studies on the dynamics of the deforming barrel shock and bow shock and the
supersonic jet-crossflow interactions [165,180,198], Kawai and Lele accompanying large-scale vortex formation from the windward jet
[186], and Wang et al. [188] performed large-eddy simulations of an boundary. The excitation of the windward jet shear layer instability by
underexpanded sonic jet injection into supersonic crossflows to obtain the turbulent crossflow was evidenced by pressure time history and
insights into key physics of the jet mixing with an emphasis on the turbulence spectra. The interaction between the turbulent structures in
impact of the unsteady turbulent boundary layer. The simulation the upstream incoming boundary layer and the jet enhanced the insta­
replicated the experimental tests of a sonic air jet injected into a Mach bility of the windward jet shear layer, which supported a more rapid

12
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 10. Representative time-series snapshots of jet fluid overlapped with divergence of velocity contours at midline plane z/D = 0 [186]. Nondimensional
time, t = t × U∞/D.

breakdown in the jet shear layer structure to the turbulent state. supersonic combustion using the wall-modeled LES method incorpo­
In supersonic reacting flows, jet flame stabilization processes in su­ rated with a novel evolution-variable manifold framework [200]. The
personic crossflows, including the fuel ignition, flame structure and simulated hydrogen-air combustion configuration was consistent with
near-wall burning phenomenon, are closely dependent on the jet- the experimental conditions in Gamba and Mungal [146]. In premixed
crossflow interactions and the associated mixing process, which im­ combustion, the Karlovitz number (Ka) characterizes the length scale of
pacts the local mixture stoichiometry, concentration gradient, local flow reaction zone relative to the length scale of the smallest dissipative
temperature and residence time. Especially at high-speed conditions, turbulent eddies (Kolmogorov length scale). Fig. 12 shows the estimated
radical formation and ignition delay are important for combustion sta­ Karlovitz number of the near-wall turbulent combustion on the center­
bilization. Gamba and Mungal [146] studied the ignition, flame struc­ line of the hydrogen jet flame. The blanked regions include the regions
ture, and near-wall burning of transverse hydrogen jets in a Mach 2.4 of non-premixed combustion at the leading edge of the jet, the regions
airstream of an expansion tube facility. The high-enthalpy freestream where the equivalence ratio ϕ is less than 0.5 or greater than 2.0, and the
had a static temperature of 1400 K (with a corresponding stagnation regions where the heat release is less than 4% of the maximum heat
temperature of 3000 K) that was above the fuel autoignition tempera­ release. The numerical results indicated that there existed a substantial
ture. Fig. 11 shows the instantaneous flame structures visualized using fraction of the flowfield where Ka > 100, which shows that combustion
OH-PLIF imaging. The jet flame stabilized in the upstream recirculation in those regions occurred in the distributed reaction zone regime. This
region by autoignition due to the sudden temperature increase after the region was experimentally identified to be dominated by distributed
bow shock. In the wake close to the wall within the boundary layer, the near-wall burning [146], and the LES analysis indicated that the
flame was stabilized by autoignition or flame propagation after partial distributed reaction zone was produced due to the large levels of tur­
premixing in the near field. The jet ignition characteristics were strongly bulence from the counter-rotating vortex pair.
affected by the jet-to-crossflow momentum flux ratio J, in which the In summary, mixing and combustion processes of jet-crossflow in­
flame was lifted and stabilized in the wake region at low J, and then teractions in the supersonic regime are dominated by the production and
moved upstream of the injection point in the boundary layer at high J. evolution of large-scale coherent structures and counter-rotating vortex
The primary source of combustion heat release was the near wall region, pairs. The evolution of coherent structures at the jet-shear layer inter­
though flames were observed at both the shear layer and the near wall face is highly unsteady and correlated to the tilting and stretching of
region at a sufficiently large value of J. It was concluded that the near eddies. The large-scale structures in the far field dissipate and translate
wall combustion could be the dominant mechanism for jet flame stabi­ in the streamwise direction close to the freestream velocity which leads
lization in supersonic flows. Jet flame stabilization in supersonic cross­ to mixing-controlled combustion. The diffusive jet flame behaviors are
flows was characterized by thin reaction layers and distributed reaction impacted by parameters such as the freestream Mach number, boundary
regions, which indicated the coexistence of multi-region combustion layer thickness, jet-to-crossflow momentum flux ratio, pressure and
stabilization from non-premixed flamelets to partially premixed density ratios, and compressibility effect. Diffusive combustion,
distributed reaction zones. including the partially premixed reaction zones and non-premixed
Candler et al. [189,199] simulated autoignition-dominated flamelets, stabilizes in the flow regions near the wall and the jet-shear

13
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 11. Instantaneous OH-PLIF images of transverse hydrogen jet (J = 5) flame at the mid-plane (top) and the wall-parallel planes [146]. The dimension is scaled by
the diameter of jet orifice d.

Fig. 12. Estimated Karlovitz number on the centerline plane of the hydrogen jet combustion (J = 5) [189]. The black line indicates the stoichiometric condition.

layer interface, in which the flow residence time, static temperature, and total pressure loss from shockwaves is large compared to the gain from
stoichiometric condition are favorable to combustion stabilization. the fast near field mixing at high Mach number regimes, which indicates
a deterioration of overall engine performance. The total pressure loss
2.1.2. Oblique jet injection from the strong jet-induced bow shock can be mitigated using an oblique
Scramjet thrust is globally determined by combustion, which is jet injection. The injector direction plays an important role in the bow
controlled by the near and far field mixing and losses from shocks, shock strength, with the shock created by an oblique injector being
frictional drag, and pressure drag [201]. For transverse jet injection, the substantially weaker than that created by a transverse injector [98].

14
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Oblique injection has been used to reduce the separation upstream of the and transverse sonic ethylene jets in a Mach 2 airstream using the
jet and to increase the jet penetration height. In addition, the structural Raman scattering method. Ethylene was distributed at both sides of the
organization at different injection angles can have a direct impact on its main fuel plume near the tunnel floor only in the case of transverse
mixing characteristics due to the large-scale structures formed at the injection. The transverse jet had a higher combustible mixing region
interfaces between the freestream and injectant fluids in the near field enclosed by the stoichiometric line and the plume boundary. The obli­
and the counter-rotating vortex pairs in the far field. The wall-parallel que injector delivered smaller fuel penetration, lateral fuel spreading,
component of the oblique jet momentum along the streamwise direc­ and degree of fuel-air mixing in the near field. Nevertheless, a unified
tion can produce a sizeable fraction of total thrust from the injection correlation of fuel plume penetration height was obtained for transverse
itself [202]. and obliques jets by using the measured ethylene mole fraction along the
Effects of oblique jet injection angle on the overall characteristics of fuel plume centerline at various freestream locations,
jet penetration trajectory and averaged concentration distribution were
extensively studied in non-reacting configurations in the 1960s and h0.01 / d0 = 1.16q0.72 (x/d0 )0.32 θ0.11 , (3)
1970s. McClinton [203] studied the effect of injection angle on the jet
where h0.01 is the measured fuel plume penetration height indicated by
penetration, mixing rate, and total pressure recovery in a Mach 4 free­
the ethylene mole fraction of 0.01 interpolated from the quantitative
stream. The hydrogen jet was injected into the supersonic crossflow at
measurements, x is the streamwise distance from the jet exit, d0 is the
angles ranging from 30◦ to 90◦ . The results for hydrogen concentration,
injector diameter, q is the jet momentum, and θ is the injector angle to
pitot pressure, and static pressure at multiple locations downstream of
the flow. Fig. 13 shows reasonably good agreement between the
the injector showed that at lower injection angles, less freestream mo­
measured penetration heights and correlation-predicted values.
mentum loss was produced, jet penetration increased downstream, and
In summary, oblique jet injection can alleviate the total pressure
mixing of the injected gas with the freestream occurred faster. Mays
losses but degrade the jet-crossflow mixing efficiency in the near field
et al. [202] investigated helium jets in a supersonic airflow at Mach 3
compared to transverse jet injection. Studies on performance deviation
using circular sonic orifices angled downstream at 15◦ and 30◦ . The
between oblique and transverse jets are mostly focused on the mixing
jet-induced bow shock close to the jet induced a small or negligible
characteristics and total pressure loss in non-reacting supersonic flows.
separation region upstream of the jet, and the bow shock angle increased
Though the oblique jet injection scheme has been used to mitigate shock
a few degrees when the jet injection angle was doubled. The increase in
losses in scramjet fueling schemes, literature on experimental charac­
injection angle resulted in a faster mixing rate in the near field, but the
terization of flame ignition and stabilization regarding oblique injectors
peak injectant concentrations and the decay of the maximum concen­
is limited for reacting high-enthalpy freestream cases. The oblique jet
tration with downstream distance were not impacted by the injection
can reduce shock losses while it suffers from reduced mixing efficiency.
angle in the far field. Schetz et al. [204] summarized the transverse and
The overall contribution of the oblique jet to scramjet performance re­
oblique mixing data from a series of jet penetration studies. The pre­
mains unclear, and the optimum injection angle needs to be further
sented results led to the conclusion that increasing the injection angle of
studied.
oblique jets could promote jet-crossflow mixing in the near field, in­
crease the total pressure loss, but contribute negligible mixing
2.1.3. Parallel jet injection
enhancement in the far downstream field.
A parallel jet in supersonic flows injects fuel through a slot or an
The reduced generation of streamwise vortices at the jet-freestream
array of jet orifices in a step/strut facing in the downstream direction.
interface was responsible for the reduced mixing efficiency in the near
Similar to oblique jet injection, parallel jet injection can further reduce
field of an oblique injector relative to a transverse injector. Gruber et al.
airstream total pressure loss and direct all the fuel injection momentum
[86,98] visualized the instantaneous flow structures of jet penetration at
to thrust generation [202]. Compressibility effects on shear layer
90◦ and 15◦ inclination in a Mach 2 freestream using Mie scattering. The
large-scale vortices developing at the jet-freestream interface were
observed for transverse injection, while these vortical features were not
observed along the interface in the near-field region of the oblique in­
jection. The oblique injector had a reduced mixing efficiency in the near
field relative to a transverse injector. McCann and Bowersox [205]
examined the mean and turbulent flowfield of a Mach 1.8 air jet injected
into a Mach 2.9 airstream at a 25◦ inclination angle. It was found that the
mean and turbulent structure (turbulence intensity, vorticity, and Rey­
nolds shear stress) were strongly influenced by two counter-rotating
vortices within the core of the injection plume.
Boundary layer thickness and yaw angle were found to also affect the
mixing characters of an oblique jet. Fuller et al. [206,207] explored an
oblique helium jet in Mach 3 and 6 supersonic crossflows. The trends of
core penetration growth rate and the maximum concentration decay
rate with respect to the injectant underexpansion pressure ratio were
different for Mach 3 and 6 freestreams under the same jet configuration.
The thinner boundary layer thickness at Mach 6 experiments was
responsible for the variance in the trend. The jet penetrated into the
boundary layer first, then swept downstream much faster outside the
boundary layer, which led to a reduced time for mixing. The injector
yaw angle did not increase the rate of decay of the maximum concen­
tration, but caused an increase in the overall injectant plume cross
section and the size of the mixing region.
Excluding the influence of the wall boundary layer, the properties of
fuel plume structure, such as shape, size, and concentration profiles, are
scalable in terms of the injector size and the injection angle. Lin et al. Fig. 13. Scalability of ethylene penetration height between transverse and
[168,169] characterized the distribution of fuel mole fraction in oblique oblique jets [168,169].

15
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

development in high speed mixing and combustion were investigated combustors are adapted from jet mixing studies of low Mach number jets
using parallel jet injection. The increased compressibility modifies the injected into quiescent flows [215–218] and coaxial/parallel flows
turbulent structure and suppresses the transverse momentum transport [219–221]. Non-circular injectors benefit from large-scale mixing at the
and the mixing layer growth rate [208]. Flow penetration does not exist flat sides and small-scale mixing due to corner flows. Coherent struc­
for tangential injection between the jet and freestream at comparable tures originating from the flat sides and corners have different spreading
static pressures, and the flow mixing in the compressible shear layer is rates and vortex deformations, which result in the complex vortex
extremely slow. In scramjet applications, the required combustor length interaction along with the evolution of streamwise vortices [211]. The
for fuel-air mixing using a parallel fueling scheme can be extremely long, upstream separation region encountered in circular injectors results in
which complicates the combustor structural design. Its dimensional increased pressure and tremendous local heat loading on the wall,
limitations in fuel penetration, mixing, and flameholding characteristics though it may ignite and/or stabilize flames in supersonic flows. The
have been identified and various mixing-enhancement strategies for flush-walled orifice injection without upstream separation is necessary
diffusive combustion configurations have been proposed and to alleviate the local thermal loading on combustor walls in the trans­
investigated. verse injection configuration. 3D injectors with their sharp corner ori­
ented toward the oncoming flow can eliminate the upstream separation
2.2. Flow mixing enhancement when the apex angles are less than a certain degree [222,223]. Modifi­
cations on the aspect ratio of injector geometry and alignment (injection
Fuel injection needs to achieve rapid macro-scale fuel-air mixing, angle and yaw angle) can improve penetration and mixing of 3D in­
permit sufficient residence time for fuel ignition, destroy large-scale jectors due to the dependence of jet-crossflow interaction on the injector
turbulence structures, and promote small-scale turbulence generation geometry and direction. Fig. 14 shows several examples of 3D injectors
for micro-mixing at a rate consistent with chemical reaction times when considered for mixing enhancement in scramjet combustors. Table 2
sufficient reactants have been macro-mixed [209]. Mixing characteris­ summarizes the studies of 3D injectors in supersonic crossflows. Other
tics in subsonic and lower-regime supersonic flows have been summa­ injectors using effects from nozzle wall asymmetry [224], nozzle exit
rized in Refs. [209–212] decades ago. Mixing enhancement in high attachment [225,226], and tabs [227,228] are rarely used in scramjet
Mach number flows has exceeded the conventional studies of particular applications and hence not included.
method applied to subsonic free jets and the two-dimensional (2D) Mixing enhancement using 3D injectors in supersonic crossflows was
shear-driven mixing due to the limitation of their effective ranges [213]. initially conceived by studies in mitigation of jet-induced boundary
With the increase of combustor Mach number, fuel-air mixing that can layer separation. A fuel injector with a sharp leading angle was origi­
be achieved through natural convective and diffusive processes in nally proposed to mitigate the 3D boundary layer separation occurring
high-speed combustion is diminished due to the flow compressibility upstream of the circular injection port in a transverse injection scheme.
effects, which significantly suppresses transverse Reynolds normal Masyakin [222] observed no boundary layer separation at the windward
stresses for momentum transport and primary Reynolds stresses for side of the triangular-shaped jet when the jet apex was oriented toward
mixing layer growth with an increasing convention Mach number [208]. the freestream and the apex half-angle was less than 12◦ . The absence of
The inherent low growth rate of supersonic shear layers pose a tech­ an upstream separation zone significantly reduced local thermal loading
nological challenge of mixing enhancement in high-Mach number flows on the combustor wall. This experiment inspired continuous research
[211]. Enhancement of mixing and associated combustor length efforts on 3D injectors that could potentially lead to better fuel mixing
reduction in scramjet combustion systems has become essential in suc­ than that through a circular orifice. Barber et al. [223] used the
cessful aircraft design [214]. Parameters considered in optimization of wedge-shaped wall orifice to deliver the injectant into Mach 3 transverse
mixing enhancement include the global distribution of fuel contours, flows at sonic speed. The wedge-shaped injection demonstrated greater
irreversibility from jet penetration and increased mixing rate, thrust penetration into the freestream, faster expansion of the jet cross-section
gain from fuel jet momentum, drag from aerodynamic friction, and in the freestream, and better mixing compared to the equivalent con­
engine’s structural efficiency. In practical designs, the variables that can ventional circular jet. The round-shaped back half of the wedge-shaped
be manipulated include the fuel jet sites and quantity, sites of mixing injector was believed to introduce flow separation in the jet wake region
devices, and the interactions between jets and mixing devices. and intensive flow disturbances. Tomioka et al. [229] modified the
In this section, promising mixing enhancement methods for scramjet wedge-shaped orifice by introducing a diamond-shaped orifice having a
applications and recent research progress are reviewed and discussed. wedge-shaped portion on both the front and back halves. At low
These methods are categorized based on physical processes involved in jet-to-crossflow momentum flux ratio, the plume from the
mixing enhancement, which include 3D-shaped jets, shock-enhanced diamond-shaped injector penetrated farther into the supersonic cross­
mixing, axial vorticity generation, and excitation of shear layer in­ flow in contrast to a comparable circular injector. The plume sharpness
stabilities. Among these methods, non-intrusive and intrusive mixing deteriorated with the increase of jet momentum, and the penetration of
designs are used. Non-intrusive designs for mixing enhancement include the plume was comparable to that from the circular injector.
3D-shaped jets (circular, elliptical, diamond, wedge, etc.) and aero­ The mixing of non-circular injectors is dominated by the secondary
dynamic configurations (shock-jet interaction, cascade jets, aero­ flow structures generated by jet-crossflow interaction. Fig. 15 shows the
dynamic ramps, pulsed detonator, flameholding cavities). Intrusive flow structures created by a sonic injection into a supersonic crossflow
designs for mixing enhancement involve insertion of geometric features through a diamond-shaped orifice [234,235]. The jet-crossflow inter­
into high-speed flows (struts, pylons, hypermixers), which produce action forms a lambda shock upstream of the injection port, causes
aerodynamic perturbance into supersonic flow by axial vorticity gen­ significant pressure rises around the front half of the injector, and gen­
eration from surface geometry and shock-induced baroclinic effects. erates a low-pressure recirculation zone downstream of the injector port.
Although they create more aerodynamic disturbances with an inevitable The barrel shock formed by the diamond-shaped injector has a stream­
increase in total pressure loss and drag, intrusive methods can effec­ lined, sharp leading edge and a nabla-shaped trailing edge due to the
tively enhance the combustion process and significantly reduce the axis switching [229], which drives the shock to expand more in the
overall length of the scramjet combustor. Active cooling and the asso­ lateral direction than in the axial direction. The diamond-shaped
ciated complexities are usually required for intrusive devices due to the injector produces two pairs of vortices at the leading and trailing
high thermal loading. edges of the port due to the vorticity generation in the corner flow. As
the injector fluid expands, two vortices undergo lateral separation along
2.2.1. Three-dimensional jets its flowpath. A lateral counter-rotating vortex pair (LCVP) is generated
Three-dimensional jets used for mixing enhancement in scramjet in the low-pressure region immediately downstream of the barrel shock,

16
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 14. Schematics of 3D injectors for mixing enhancement.

Table 2
3D-shaped jet studies for scramjet combustion.
Literature Injector Injectant Mj Mf Method Study content

Tomioka et al. (1) Air 1.0 3.0 Surface oil flow, shadowgraph, pitot Effects of swept-back and yaw angles on mixing and penetration
[229–231]
Bowersox et al. (1) Air 1.0 5.0 Surface oil flow, shadowgraph, Mie Effects of incidence angle and jet pressure on flowfield
[232,233] scattering, pressure-sensitive paint, pitot documentation, jet penetration, shock-induced total pressure loss
Srinivasan et al. (1) Air 1.0 2.0,5.0 RANS, DES Secondary flow structures
[234,235]
Kobayashi et al. (1) Air/ 1.0 2.0 RANS, NO-PLIF, OH-PLIF Effects of igniter torch flow on the near-field flow structure
[236] C2H4
Grossman et al. (1) He 1.0 4.0 gas sampling Effects of jet momentum, incidence angle, yaw angle on mixing
[237] and total pressure loss
Gruber et al. [98, (2) He, CO2 1.0 2.0 Mie Scattering Penetration and mixing characteristics
238]
Wang et al. [239] (2) Air 1.0 1.6 LES Shock/jet interaction, jet shear layer vortices and their evolution
Foster et al. [240] (1)(3) O2 0.3 1.2 RANS Performance and flow behavior of various injector geometry
Ogawa et al. (1)(3)(5) H2 1.0 5.0 RANS Effects of orifice shapes on fuel mixing characteristic
[241–243]
Barber et al. [223] (4) He 1.0 3.0 Surface oil flow, shadowgraph, gas Upstream separation, jet penetration and mixing
sampling
Hirano et al. [244] (1)(3)(5) He 1.0 2.5 RANS and experimental gas sampling Effect of injector geometry on penetration and mixing
(6)(7)
Hariharan et al. (1)(4)(10) He 1.0 3.0 RANS Effect of injector geometry on mixing, separation, and total
[245] pressure loss
Axdahl et al. (1)(5)(7) H2 1.0 7.67 RANS Effect of fuel injection strategies on mixing efficiency and stream
[246] thrust potential
Kouchi et al. (7) H2 1.0 2.44 Gas sampling Combustion enhancement by stinger injector
[247]

which entrains the incoming freestream boundary layer fluid into the the axial counter-rotating vortex pair as they move past the barrel shock.
injector fluid. A pair of cone-shaped structures is formed at the down­ Kobayashi et al. [236] characterized the flow structures and the mixing
stream edge of the vortex pair. The entrained fluid is ejected from the characteristics of a flush-walled transverse jet from a diamond-shaped
cone-shaped structure in the downstream direction and transported by injector in Mach 2 airflow, as shown in Fig. 16. In the near field of the

17
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

the decreasing half angle of the wedge-shaped injector. There exists a


strong correlation between the aspect ratio and the penetration height,
and the reduction of the injector width reduced the flow disturbance
indicated by the maximum static pressure on the injection wall surface,
as shown in Fig. 18. The stinger-shaped injector was shown to have the
most effective penetration performance in conditions of low dynamic
pressure ratios (J < 3) relative to a circular injector. Kouchi et al. [247,
249] compared the combustion performance of stinger and circular
shaped injectors in a Mach 2.5 airstream with a stagnation temperature
of 2060 K. In the low fuel-equivalence ratio regime (ϕ < 0.6), the stinger
injector produced 10% higher pressure thrust than the circular injector
due to the high jet penetration. Further increasing the fueling rate led to
the generation of a pre-combustion shockwave in front of the injector,
which decreased the momentum flux of the crossflow and diminished
the advantage of the stinger injector.
Ogawa [241,242] numerically compared the mixing performance of
circular, triangular, diamond, square, and rectangular-shaped injectors
in terms of the mixing efficiency, total pressure recovery, fuel penetra­
tion, and streamwise circulation at various injection angles, aspect ra­
tios, and jet momentums. The leading-edge geometry of the injector had
primary influence on the 3D bow shock strength, whereas the aft shape
of the orifice influenced the fuel mass flow rate injected into the rear of
the jet plume. The triangular jet showed the strongest streamwise cir­
culation. The longitudinally slender injector shapes, including diamond
and rectangular injectors, enhanced the streamwise vorticity generation
and fuel penetration due to the buffering (shielding) effect. All of the
shaped injectors were found to incur approximately the same level of
total pressure loss at the end of injection, except for the slot (slender
rectangular) injectors which exhibited slightly smaller losses.
While sharp leading edge orifices were introduced to improve per­
Fig. 15. Flow structures of a sonic diamond-shaped jet in a Mach 5 flow [234].
formance at low jet momentum ratios, elliptical orifices were introduced
as a low-disturbance injector at high jet-to-crossflow momentum flux
jet penetration, the jet from a diamond injector has a wider area but less ratios in supersonic crossflows [229]. Elliptic jets with small aspect ra­
penetration depth than that from the circular injector case. The jet tios produce an asymmetric distribution of boundary layer momentum
plume was highly elongated in the lateral direction, and it had pronged thickness around the circumference of the jet orifice, which results in the
structures in the center observed only in the diamond injector. In the asymmetric vortices shedding from the nozzle exit [238]. Variances of
further downstream region, the counter-rotating vortex pair dominated spreading characteristics of the jet plumes in minor and major axis
the mixing process as evidenced by the far-field cardioid shape for both planes lead to the axis switch, i.e., the jet spreading of the minor axis
injectors. plane surpasses that of the major axis downstream along the flowpath.
Mixing performance of the diamond-shaped orifice is affected by the The different spreading features between elliptical jets and circular jets
jet alignment to the supersonic crossflow. Low downstream angled in­ have been revealed to affect the convection of the shear layer eddies, the
jection of the diamond-shaped orifice was studied by Tomioka et al. characteristics of the upstream bow shock, and separation of the up­
[229–231] and Grossman et al. [237] in Mach 3 and Mach 4 airflows, stream flow.
respectively. Angled injection was observed to enhance the penetration Compared to circular jets, elliptic jets enable faster lateral spreading
depth and jet plume width while reducing the total pressure losses at an and reduced jet penetration in the near field, which are affected by the
increased jet-to-crossflow momentum flux ratio. Bowersox et al. [232, jet aspect ratio and injection angle. Gruber et al. [98,238] examined the
233,248] compared the jet penetration performance between diamond sonic transverse injection from elliptic and circular nozzles into a Mach
and circular orifices with various incidence angles in a Mach 5 free­ 2 supersonic crossflow at J = 2.9 using planar Rayleigh/Mie scattering.
stream. Fig. 17 presents the overall penetration correlation of diamond Both jets induced 3D bow shocks and large-scale eddies at the interfacial
and circular jets. In the correlation equations, x and y are the axial and regions between the jet and freestream. The elliptic jet spread more
transverse coordinates of jet trajectory at its centerline, d is the jet rapidly into the crossflow in the lateral direction than the circular jet
equivalent diameter based on area, and Jeff is the transverse component while suffering a 20% transverse penetration reduction. At the earliest
of jet-to-crossflow momentum flux ratio J. Additionally, Fuller et al. stage of the interaction, the lateral spread of the elliptic and circular jets
[206,207] applied a low sweepback angle and a yaw angle to the was nearly equal. The counter-rotating vortices emerged and moved
diamond-shaped injector. Increasing the yaw angle of the injector further apart laterally along the flowpath. The elliptic jet had an
increased the overall plume cross-sectional area but decreased core increased separation between the counter-rotating vortices, and hence
penetration into the freestream. Tomioka et al. [229] also found that enhanced central fluid entrainment into the jet. The flow region with
increasing the yaw angle configuration could reduce the maximum large counter-rotating vortices was shown to be a critical location
mixing concentration. providing efficient freestream fluid entrainment into the jet core. Ogawa
Hirano et al. [244] analyzed the disturbance of the jet to the main­ [243] numerically confirmed that the streamwise aspect ratio of the
stream and the mixing effectiveness of multiple injectors including elliptic injector played a major role in the mixing efficiency and a sec­
triangular, arrowhead, diamond, rectangular, and circular orifices. A ondary role in fuel penetration. Low-angle fuel injection through a
sonic helium jet was injected into a Mach 3 airstream for all injector highly elliptic orifice with wide spanwise spacing demonstrated the
geometries. Both experimental and numerical results indicated that or­ most comprehensive advantages in terms of the fuel penetration, total
ifices having sharp leading edges mixed more effectively than the case of pressure loss, streamwise vorticity, and mixing efficiency.
blunt leading-edge injectors. The penetration height was increased with Aside from general deviations of jet spreading features between

18
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 16. NO-PLIF images of diamond and circular orifices in lateral view [236].

circular and elliptic injectors, the jet shear layer vortices demonstrate penetration compared to a circular injector. The frequency relevant to
distinct dynamics during downstream convection. Wang et al. [239] the eddy formation and the spacing between the core of the jet shear
simulated circular and elliptic sonic jets injected into a Mach 1.6 layer vortices was extracted from the instantaneous vortical structures.
crossflows at a jet-to-crossflow momentum flux ratio of J = 1.7 using The formation frequency, defined as St = f∙D/U∞, was 0.469 for circular
LES. Fig. 19 compares instantaneous flow structures induced by circular injection and 0.625 for elliptic injection, where f is vortex shedding
and elliptical jets. The elliptical jet induced a weaker bow shock, miti­ frequency, D is jet diameter, and U∞ is freestream velocity. At the
gated flow separation attached to the wall upstream of the jet, increased streamwise location x/D = 2.0, the vortices were spaced at 0.4D-3.2D for
spreading rate in the spanwise direction, and reduced the transverse jet the circular jet and 0.4D-1.8D for the elliptic jet.

19
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

uncertain. Additionally, jet flame dynamics have yet to be temporally


and spatially resolved using state-of-art high-speed laser diagnostics in
these high speed flow conditions.

2.2.2. Shock-enhanced mixing


Introducing incident shockwave can promote supersonic fuel-air
mixing. Mixing enhancement by shock-induced interactions could be
an effective and economical approach because shockwaves are inherent
flow structures generated in supersonic flowpaths [251,252].
Shock-enhanced mixing involves the perturbation of the shear layer
formed between the airflow and injected fuel by an incident shock in
order to enhance the mixing downstream of the shock/shear-layer
interaction region. Shockwaves in hypersonic propulsion systems
include (1) incident and reflected shocks, (2) shocks formed by flow
separation regions, (3) shocks generated by jet injection (fueling and
cooling) and fueling struts, (4) interference shocks generated in corner
flows, intersection regions, and shock-shock interactions, (5) flowpath
components (step/cavity flameholder), and (6) active perturbation
sources (pulsed jets, pulsed detonator, laser-induced spark) [253].
Shock enhancement of supersonic flow mixing was originally proposed
by Marble et al. [254] based on the baroclinic effect. The rate of change
of vorticity ω, is derived by taking the curl of the momentum equation
and can be written in the following form [211]:
Fig. 17. Penetration correlations of sonic diamond-shaped/circular jets in a Dω ∇p × ∇ρ
Mach 5 flow [233]. =− + (ω ⋅ ∇)u − ω(∇ ⋅ u) + ν∇2 ω + (∇ν) × ∇2 u (4)
Dt ρ2

Mixing performances of an elliptic nozzle and a circular nozzle with where ρ is the density, p is the pressure, u is the velocity, and ν is the
vortex generating tabs were also experimentally evaluated in parallel kinematic viscosity. The first term on the right-hand side of the equation
injection into a Mach 2 freestream [250]. The total mixing evaluation is the baroclinic torque, which exists when the pressure and density
method evaluates the mixing potential of the flowfield by employing the gradients are not aligned. The shock-induced vorticity was observed in
large-scale mixing parameters such as perimeter, area, and shape factor. the configuration of a planar shock passing over a circular region of light
Experimental results showed that the circular-with-tabs nozzle provides gas enclosed by air [254,255]. The imposed pressure gradient at the gas
the most desirable large- and small-scale mixing enhancement interface with different densities produces vorticities, which lead to
characteristics. rapid gas interface distortion and coalesce into a counter-rotating vortex
Although extensive studies exist on the mixing characteristics of 3D pair for large-scale rapid convective mixing. Additionally, shock/­
jets, the performances of those 3D jets in reacting flows under the impact boundary layer interactions amplify the incident turbulence across the
of combustion heat addition and flame anchoring have yet to be further shockwave [256]. Shock-turbulence interaction involves generation of
investigated. For instance, the recirculation zone upstream of a circular velocity fluctuations from incident acoustic and entropy fluctuations,
jet is considered to be an ignition source for jet in a supersonic crossflow, intensification of small scale motions by shock distortions, and forma­
though it exacerbates the heat loading on the combustor wall, while the tion of local pressure gradients and streamline curvature for additional
upstream flow separation can be avoided for 3D jets with a sharp leading destabilization; they result in intensified fluctuation energies from the
edge. Effects of the upstream anchored flame on the flame stabilization mean field through shock motion, and the direct amplification of inci­
and overall combustion enhancement are still unclear. In high-enthalpy dent dynamic vorticity [253]. These shock-induced fluid dynamic ef­
flows, the flame may always be stabilized at the leeward side of the 3D fects have immediate significance to mixing enhancement downstream
jets, which can exacerbate heat loading on the wall similar to the up­ of the shock impingement region in supersonic flows.
stream flow separation. Whether it is necessary to alleviate flow sepa­ An incident shock impingement on supersonic shear layers is
ration upstream of the fuel jet as a part of fuel injector design remains

Fig. 18. Correlations between the jet penetration depth at x/d = 21.8 and (a) the aspect ratio of injectors, (b) the maximum wall pressure values [244].

20
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 19. Density gradient of circular (top) and of elliptic (bottom) injections [239].

encountered in dual-combustion ramjet (DCR) engines. The DCR engine combustor while the major portion of the air passing over the cowl
concept was proposed for hypersonic volume-limited missiles in the surface enters the isolator and then supersonic combustor. The subsonic
Applied Physics Laboratory (APL) at Johns Hopkins University in the combustor may burn part or all of the fuel in order to match the required
1980s [257]. Fig. 20 shows the schematics of a DCR engine and the operating conditions. Hot products from the subsonic combustion are
shock-mixing layer interaction after a circumferential splitter plate. Air injected into the main supersonic combustor to pilot and stabilize the
is initially compressed by an axisymmetric external compression surface combustion downstream. The hot mixture from the subsonic combustor
and then divided by an internal cowl into inner and outer flow ducts. A and the air from the isolator merge and react in the wake of the rim of
smaller fraction of air passing over the inner duct enters the subsonic the subsonic combustor exit, which is equivalent to a supersonic mixing

Fig. 20. Schematics of (a) a DCR engine concept with associated shock structures [257], and (b) the shock-mixing layer interaction [258].

21
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

layer developing after the flow separating at the splitter plate [252]. An
incident shock generated by a wedge/bump impinges on the supersonic
shear layer and produces shock refraction and reflection between the
mixing layer and the wall. The dump combustor can act as either a pilot
or a gas generator to assure that heat can be efficiently released in the
supersonic combustor. Fuel is added to both streams for its operation as
a pilot, whereas all of the fuel is added within the dump combustor for its
operation as a gas generator, in which the main combustor becomes a
supersonic “afterburner”.
In strut-based scramjet combustors, shockwaves generated at the
strut leading edge and subsequent reflection on the combustor wall
interact with the shear layer after the strut. In these configurations, axial
vorticity generators are usually integrated with the strut body to
enhance the fuel-air mixing, in which the shock impingement on the
mixing layers is a secondary factor to increase the mixing efficiency.
Mixing enhancement by axial vorticity generation will be reviewed in
the next section.
Mixing enhancement was experimentally/numerically observed on
supersonic shear layers after a flat plat impinged by an incident shock. In
a non-reacting flow configuration, a sonic helium jet downstream of a
rearward-facing step was mixed with a Mach 2.5 nitrogen stream to
simulate the mixing region at the rim of the DCR subsonic combustor
and the vicinity of a scramjet flameholder [258]. The mixing layer was
impinged by an oblique shock formed by a planar wedge with a
leading-edge angle of 15◦ . Significant spreading of the shear layer
occurred downstream of the shock/shear-layer interaction region. The
result confirmed the possibility to enhance mixing and hence combustor
efficiency by shock impingement, although the total pressure losses
associated with the shock generation and the shock/shear-layer inter­
action could be increased. Nevertheless, the spreading mechanism and
the extent of increase in mixing efficiency in terms of concentration and
density variation were not determined in experimental results. Fig. 21. Spanwise vorticity for three cases: (a) the shock-free mixing layer,
Flame stability can be enhanced by effects introduced through shock incident shock impingement at angles of (b) 20.21◦ and (c) 25.62◦ [260].
impingement in reacting flows. Shockwave-induced effects on mixing
and flame stability limits of a sonic hydrogen jet flame stabilized on the plates with various thickness [252]. The recirculating wake flow behind
axis of a Mach 2.5 flow was experimentally investigated [259]. In this the splitter plate was crucial in determining the mixing behaviors of the
configuration, the thick-lipped fuel tube was used as a bluff body. coflowing fuel and air. In the near field, the primary mixing layer started
Experimental results showed that the flame lengths were decreased by immediately downstream of the recirculation zone and two oblique
20% due to the enhanced fuel-air mixing when an optimum shock shocks stemmed from the edge of the splitter plate. Increasing the
location and shock strength were chosen. When the shock interacted thickness of the splitter plate resulted in shear layer expansion and
with the flameholding recirculation zone at the rim of the thick fuel intensified interactions with shocks, which created broader low-speed
tube, the shock-induced adverse pressure gradient elongated the recir­ and high-temperature regions for flame ignition. The flame was
culation zone, which substantially improved the flame blowout limit. attached to the wake flow for a thin splitter plate, whereas the flame was
Shock-induced hairpin vortices from the baroclinic effect are lifted for a thicker plate due to the strong flow expansion with increasing
responsible for the increased growth rate of the supersonic mixing layer. plate thickness.
In the shock-mixing layer region, hairpin vortices were produced from Additionally, perturbations on initial flow conditions upstream of the
the interaction of the incident shocks through the baroclinic effect as shear layer origin were found to affect the spreading rate of supersonic
shown in Fig. 21 [260]. The increase of mean vorticity was mainly shear layer more than the shock-induced hairpin vortices in the planar
caused by the streamwise velocity gradient along the transverse direc­ shock-shear layer interaction. Effects of shock impingement on a
tion. The vorticity thickness of the shocked-mixing layer experienced a boundary layer in a Mach 5 supersonic freestream were experimentally
sudden decrease in the vicinity of the shock impingement point followed investigated [263]. The incident shock produced by a 10◦ planar wedge
by a more rapid growth than that of the shock-free mixing layer due to impinged on the shear layer formed by a boundary layer leaving the
the induced hairpin vortices. The streamwise Reynolds stress and the trailing edge of a flat plate. An internal 2D Mach 3 nozzle was installed
turbulent kinetic energy of the supersonic mixing layer were modified to the flat plate. The shear layer thickness after the flat plate was
by the hairpin vortices and the subsequent vortices, which separated correlated with the difference in thickness of the boundary layer at the
from the hairpin vortices in the transient stage of the mixing layer. The shear layer origin. Perturbation on the initial condition upstream of the
transportation and convection of turbulent kinetic energy were signifi­ shear layer origin caused a significant increase in the spreading rate near
cant at the outer regions of the mixing layer after shock impingement, the plate’s trailing edge, but the growth rate returned to the undisturbed
which indicated the enhanced energy transport and convection between value rapidly. Nevertheless, the shock incidence on the shear layer had
the mixing layer and the mainstream by incident shocks. Additionally, no significant effect on the local thickness or spreading rate near the
the mixing layer interacting with an oblique shock can lead to diffusion shock. Shear layer structure represented by the cross-correlations of
flame ignition at an intermediate location across the mixing layer. Such fluctuating pitot pressures was not modified by shock impingement. In a
phenomena were studied analytically in Refs. [261,262]. three-dimensional configuration, a shock emanating from a bump in the
The finite thickness of the splitter plate is an important factor that sidewall of the combustor traveled outward into the flow and across the
affects vortex evolutions in the shock-shear layer interaction region. shear layer downstream of the splitter tip [264]. Flow visualization
Fig. 22 shows the flowfield characteristics in the near field of splitter using planar Mie scattering showed that co-rotating or counter-rotating

22
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

total pressure loss. Additionally, the shock-jet interaction modifies the


size of the recirculation zone downstream of the jet, increases the
airstream entrainment into the recirculation zone, elongates the char­
acteristic flow residence time in the recirculation zone, and extends the
combustible region.
Impact of oblique shocks on transverse circular jet in a Mach 2 fa­
cility was observed in experiments [268]. A leading bevel was used to
turn the flow by either 5◦ or 7◦ . Experimental and numerical data
showed that there was a fairly large shift in mixedness between the case
without shock and the one with a 5◦ shock generator. The 7◦ shock
generation angle had marginal effect on the mixedness but introduced a
large drop in total pressure relative to the 5◦ deflection case, which
illustrated the downside of using shocks to increase mixing within a
scramjet. Nevertheless, the study provided no criterion on the optimal
incident shock angle, which balances the overall mixing efficiency and
total pressure loss for ideal combustion stabilization and propulsion
performance.
To better visualize the dominant effect of the incident shock inter­
acting with a sonic jet, planar 2D configurations were investigated [269,
270]. A slot jet was used to deliver a NO/N2 mixture transversely to the
Mach 2.5 crossflow. When the incident shock was introduced down­
stream of the injection slot, flameholding was possible at a lower total
temperature of the supersonic airstream. Fig. 23 shows the flowfield
configuration and the flameholding regime in terms of the total tem­
perature and the incident shock location. Recirculation downstream of
the injection slot visualized by NO-PLIF and PIV was enlarged after
shock impingement. The fluorescence intensity was increased in the
recirculation zone with the incident shock impingement, which indi­
cated that the entrainment of the airstream into the recirculation zone
was enhanced by shock-jet interaction. The velocity vector diagram
parallel to the wall showed a large-scale circular flow formed behind the
injection slot. The 3D flow dynamics of the large-scale recirculation was
responsible for enhancing the entrainment and mixing as well as
extending the residence time of reactants in the recirculation zone. This
study was focused on the shock-induced effects on the recirculation zone
at the leeward side of a jet, but no conclusion was made on the impact of
the shock on the production of large-scale coherent structures at the
windward side of the jet.
An oscillating shock was considered to be a particularly efficient
method in shock-shear layer interaction for mixing enhancement by
generating fluctuations in energy directly from the mean flow [253].
Shock oscillation can be produced by many aerodynamics/fluid dy­
Fig. 22. Vorticity and velocity vectors in the near field with three different namic mechanisms, such as active energy deposition (laser-induced
splitter-plate thicknesses [252]. spark, pulsed detonator), utilization of jet-induced unstable flowfields
(jet instabilities and forced fluidic oscillation), unsteady shock motion in
streamwise vortices were generated by the disturbance from the side­ boundary flow transporting into a mixing region (separated flow [271],
wall. Small disturbances in the supersonic nozzle sidewalls (i.e., the turbulent boundary layer [272], resonator [273,274]), and instream
origin of shear layer) lead to the generation of significant perturbation to bodies (fuel injection strut at shock detachment limit [271], the
the shear layer. Numerical results [265] showed that mixing enhance­ Hartmann-Sprenger tube [275]). Each of these mechanisms has its
ment using shockwaves could only be effective if the stimulation was optimal working condition and application range that may not be
spatially persistent and was started as early as possible in the upstream. consistent with scramjet operation conditions. For scramjet applica­
The optimized shock position needed to be determined in order to obtain tions, mixing enhancement techniques need to maintain a steady flow
the maximum overall combustor performance due to the tradeoff be­ during the entire flight envelope. Shock interaction in boundary layers
tween the enhanced mixing/combustion efficiency and the decreased inside the scramjet combustor may generate intense unsteady flow
total pressure recovery [266]. separation and shock train propagation that can reduce the effective
An incident shock interacting with a transverse fuel jet can be core flow area and potentially lead to inlet unstart. A wall cavity is
encountered at the isolator-combustor region in dual-mode scramjet usually employed for flameholding purposes even though it can enhance
engines. The incident shock interacts with the jet-induced bow shock the fuel-air mixing due to the shock generation. Fuel delivery struts are
and impinges on the wall at the leeward side of the jet. Shock commonly used in combination with axial vorticity generation to
impingement was found to reduce the jet penetration depth and reduce enhance mixing, and the intensity of shocks standing at the strut leading
the fluctuating level of the jet’s periodic motion [267]. Nevertheless, the edge is minimized to reduce the total pressure loss along the flowpath.
shock-shear layer interaction changes the production of large-scale Hartmann-Sprenger tubes are impractical in scramjet engines due to the
coherent structures at the fuel-air interface to accelerate the mixing excessive heat loading and total pressure loss. Shock oscillation methods
process, which depends on the shock intensity. There exists an optimum available for scramjet applications include pulsed jet injection,
incident shock angle and strength to enhance jet mixing and mitigate laser-induced unsteady shock, and pulse detonator interaction with a
fuel jet.

23
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 23. Schematic of the flowfield and relationship between total temperature of airstream and location of the incident shockwave at flame extinction [269].

Pulsed jet injection can increase jet penetration and mixing due to was controlled by changing the duty cycle without any increase of the
flow unsteadiness from the starting vortex ring and the subsequent mass flow rate. Pulsed injection has a higher flux in the vicinity of the
interaction of vortex rings [102,276]. The starting vortex at the pulse injection plane and a lower peak injectant flux in further downstream
cycle induces the vertical velocity to increase the vertical jet penetra­ regions compared to continuous injection. Mixing enhancement by
tion. The injection pressure is increased in comparison to a continuous pulsed injection in the far field, beyond 6–8 times the jet orifice diam­
jet due to the duration in each cycle when the jet is not injected. This eter, was minimal. Combinations of frequencies and pulse widths are
ensures that the overall jet mass flow rate and jet diameter remain also critical factors of pulsed jet injection affecting mixing enhancement
constant. The jet penetration depth can be increased by the augmented [277]. It was found that the pulse width was of more importance than
injection pressure. Penetration and mixing of a pulsed jet can be affected frequency for mixing enhancement through pulsed injection in a Mach
by jet orifices having various diameters [276]. The pulsed injection was 1.38 crossflow. Mixing augmentation increased as the pulse width
conducted at pulsation frequencies up to 1 kHz. The penetration height decreased, whereas the effect of frequency was not substantial.

Fig. 24. Numerical schlieren of shock and vortical structures at the center plane during one cycle of pulsed jet injection [279]. D is the jet diameter.

24
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Additionally, the pulsed injection upstream of a flameholding cavity


could affect its flow entrainment into the cavity [278].
Evolution of shock systems and vortical structures during one pulsed
injection cycle is shown in Fig. 24 [279]. A sonic air jet is injected
transversely into a Mach-1.6 airstream in a non-reacting configuration.
Fundamental mechanisms in terms of multi-scale vortical structures,
complex shock systems, mixing properties and turbulence behaviors
were obtained from LES simulations. During the evolution of the pulsed
jet interacting with the crossflow, the bow shock oscillated and
deformed as the compression waves propagated upward, while inter­
mittent large-scale coherent vortices were formed due to the windward
jet shear roll up. The large-scale jet shear vortices were formed along the
interface of the jet and crossflow and promoted the engulfing of the
crossflow fluid into the injectant. Turbulent kinetic energies on the
leeward side of the barrel shock and downstream of the Mach disk were
amplified due to the complicated shock-vortex interaction and the
deformation of large-scale vortices. As a result of the diffusion of the
pulsed vortical structures, the pulsed jet showed a larger spatial corre­
lation scale in comparison to the steady jet. The spectral analysis indi­
cated that the forcing pulsations took over the primary frequency of the
flowfield.
A laser-induced plasma inside a jet plume can perturb a steady jet to
simulate the pulsed jet injection for mixing enhancement in hypersonic
flows. The possibility for using steady and periodic pulsed energy
deposition to stimulate a vortex breakdown during the shock-vortex
interaction was initially demonstrated by numerical simulation [280].
The effects of a laser-induced energy pulse on the jet flowfield in a Mach
2.25 crossflow was investigated experimentally [281]. A laser-induced
optical breakdown in air was generated in the center of a sonic under­
expanded jet exit at three vertical locations, i.e., 0D, 1.5D, 3D, where D is
the jet diameter. From schlieren images and PIV data, it was observed
that the focused laser pulse generated a blast wave to disrupt the
structure of the barrel shock and Mach disk. Vortices were formed and
convected downstream in the jet-shear layer. Although the mechanism
of vortex formation was not completely understood from these experi­
ments, the results suggest the possibility of using laser excitation as a
control technique for mixing applications involving a transverse jet
flowfield. Fig. 25 shows the concept of a laser ignition scheme in a hy­
personic inlet showing the blast wave-induced mixing in a scramjet inlet
with hydrogen injection [282]. The laser formed a kernel of
Fig. 25. Laser-induced plasma for mixing enhancement and ignition in hy­
high-temperature plasma inside the fuel plume to ignite the flow. The
personic flows [282].
expanding plasma kernel generated a blast wave that collided with the
surrounding flow and led to massive disruption of the flow structures
around the jet. The discharge of hot gas into the supersonic flow affected transverse jet. A strong bow shock created by the extremely under­
by the laser produced stronger vorticity than the blast wave as it passed expanded exhaust resulted in significant flow separation and recircula­
through the fuel-air interface. Simulation results showed that the rem­ tion upstream of the PD exit. The jet plume was initially pushed outward
nants of laser-induced plasma rolled up into a powerful vortex ring and by a large barrel shock structure and a counter-rotating vortex pair
noticeably increased the fuel plume area and the volume of well mixed formed at the PD exit. The jet was entrained into the counter-rotating
reactants. vortex pair from the PD, providing deeper distribution into the super­
A pulse detonator (PD) can effectively penetrate into the core su­ sonic flow in the blowdown process. The unsteady blowdown process
personic flow for mixing enhancement and flame ignition [103]. The also caused the jet plume to fluctuate and disperse into the crossflow and
transient pressure behavior of a pulse detonator is analogous to that of PD plume. The cross-sectional area of the jet wake visualized by NO-PLIF
pulsed injection in a supersonic crossflow. However, the PD is more imaging was increased by six times due to the pulsive flow perturbation.
powerful for mixing enhancement due to its high pressure and Differences in PD plume with various upstream injection locations
high-speed plume from the detonative combustion. The PD plume pro­ indicated the existence of an optimal staging distance for maximum
vides sufficient localized blockage for substantial penetration into the mixing and penetration.
supersonic crossflow and enhanced near-field mixing due to large vortex The PD provides a radical-rich exhaust to the wake of the transverse
generation. Additionally, a pulse detonator can provide a jet, which can alter the combustion process of the upstream jet. Fig. 27
high-temperature and radical-rich exhaust to the supersonic core flow presents the evolution of hydrogen flame structures during a PD blow­
with minimal energy expenditure for combustion sustainment in down process in a high enthalpy airstream at Mach 2.35 [283,284]. The
scramjet engines. penetration depth of the instantaneous reaction zone into the crossflow
The mixing enhancing process using a pulsed detonator in a Mach 2 was increased by approximately 150% by the staged PD relative to the
airflow with a non-reacting configuration was experimentally investi­ case without a PD. The interaction of the upstream jet and PD plumes
gated [103]. The PD provided high-pressure, high-temperature, and also induced macroscopic changes in the global combustion character­
high-velocity exhaust with blowdown times of 4–8 ms. Fig. 26 shows the istics of the transverse jet. The PD exhaust regions which maintained
transient flow dynamics of the PD plume interacting with an upstream sufficiently large temperatures to support the OH radicals extended

25
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 26. Transient flow dynamics of pulse detonator plume with upstream jet injection [103].

downstream and towards the shear layer. The combustion stability along mixing enhancing methods on scramjet operation.
the shear layer was increased by the high-temperature radicals trans­
ported to the primary jet-freestream interfaces. 2.2.3. Axial vorticity generation
In general, mixing enhancement by incident shockwaves has proved Vortex enhanced mixing, often referred to as hypermixing, uses
to be effective in both non-reacting and reacting flows, though further streamwise vortices to mix two streams. Axial vorticity generation in the
investigations are necessary to understand vortex generation and mixing layer has been suggested mainly for enhancing the mixing of
transportation during the steady and unsteady shock perturbation pro­ high-speed fuel-air flows with downstream parallel fuel injection, which
cesses. In particular, it remains unclear whether the unsteady shock is preferred over normal injection due to its lower stagnation pressure
perturbation can lead to unexpected pseudo shock oscillation and hence loss and extra streamwise momentum for thrust generation [209,212,
scramjet mode transition. In hypersonic propulsion systems, pseudo 285]. Various axial vorticity generators have been designed for both
shock oscillation may be self-sustained or forced due to acoustic and parallel and angled injection configurations based on wall surface ge­
fluid dynamic perturbations from unknown sources. The exerted fre­ ometries and aerodynamic arrangements, including compressi­
quency for jet mixing at resonant conditions may exacerbate the un­ ve/expansive ramp injectors, contoured-wall injectors, aerodynamic
steady pseudo shock dynamics and finally lead to inlet unstart. Further ramps, and strut/pylon-based injectors. The mixing ability of hyper­
studies are necessary to address the potential issues caused by unsteady mixers depends on the amount of streamwise vorticity generated by a

26
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

particular hypermixing injector, the strength and scale of vorticity, and


the hypermixer’s proximity to the fuel-air interface [286]. During the
past two decades, mixing enhancement using axial vorticity generation
has been the subject of numerical simulations and laboratory experi­
ments ranging from subsonic to supersonic flows, see Refs. [131,147,
148,209–212,287,288] for reviews. These investigations generally
indicated that axial vortex generation could be effective for combustion
stabilization over a wide range of flow Mach numbers.
Compression ramp injectors shed streamwise vorticity to directly
energize the macroscale mixing. Fig. 28 shows the swept/unswept
compression surfaces used to deliver the fuel parallel into the high-speed
freestream at the end of the ramps. The oblique shock generated by the
ramp creates a high-pressure region on the top of the ramp. Flow
downstream of the local Mach waves at the ramp surface expands into
the low-pressure regions at the sides of the ramp [286]. Flow separating
at the ramp edge recirculates inside the corners formed between the
combustor floor and the vertical wall of the ramps. The streamwise
vortices generated by compression ramps are mainly formed in the flow
separation regions. The angle of the compression ramp and the addition
of sweep to the ramp edges will impact the strength of the generated
vortices and the mixing efficiency. Increasing the compression ramp
angle broadens the pressure difference at the ramp surface and at its
sides to boost the strength of the axial vortex. The sweep feature de­
creases the pressure at the floor-ramp corners due to the increased flow
expansion by swept sidewalls. Decreasing pressure exacerbates the flow
separation in regions at the side of the ramp, enhances the flow recir­
culation back toward the ramp sides, and results in intensified vorticity.
Additionally, ramp-induced compression and expansion waves impinge
on the fuel-vortex structure in the flow, which leads to the generation of
baroclinic torque on the flow and possible vortex breakdown for further
mixing enhancement [209,289].
Fig. 29 presents the general effects of ramp angle and sweep angle on
mixing efficiency of fuel injected from a ramp at the combustor exit
[290]. The mixing efficiency at a streamwise cross-section is the ratio of
the fuel mass at that cross-section that can react completely (to the
extent that it has mixed) to the fuel mass required to consume all the
available oxygen (for a fuel-rich mixture) or the fuel itself (for a fuel lean
mixture). There exists a limit to the increase of mixing efficiency
through inducing more swirl into the flow by a ramp. Nevertheless, the
generation of a stronger oblique shock from the increased ramp angle
was found to enhance mixing of the fuel injected from the wall slot.
Two-thirds of the increase in mixing was because of the effect of the
stronger shock on the slot flow and only one-third was due to increased
axial vorticity. Strong sensitivity of mixing to sweep angle was found in
the range below the Mach angle.
Combustion process associated with the compression ramp injectors
is closely coupled with the mixing dominated by the streamwise
Fig. 27. Center-plane OH-PLIF images of a PD blowdown process in a super­ vorticity and shocks/expansion waves. Hartfield et al. [291,292] con­
sonic reacting flow [283,284]. ducted non-intrusive quantitative measurements of the injectant mole

Fig. 28. Flow features of (a) unswept and (b) swept compresion ramp injectors [214,286].

27
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 29. Effects of (a) ramp angle and (b) sweep angle on mixing [290].

Fig. 30. Spatial flame structures visualized using OH-PLIF downstream of a swept ramp injector [293].

delivered into combustion-heated supersonic airflows (M∞ = 1.9) and


ignited by an oblique shock generated by a small wedge fixed at the
upper wall of the combustor. As shown in Fig. 30, flame quenching
caused by expansion behind the edge of the wedge and flame relight by
the compression shock occurred along the flowpath. The vortex-like
flame zone developed downstream of the injector spread to the com­
bustion chamber side walls. In further downstream flow regions, the
influence of vortex structures on the mixing process dissipated and
turbulent mixing started to become the dominant mixing mode.
Northam et al. [214] explored the mixing enhancement for
wall-mounted parallel injection using swept/unswept ramp injectors. A
hydrogen jet (Mj = 1.7) was injected into combustion-heated supersonic
flows (M∞ = 2) in the streamwise direction. In reacting flows, the
combustion efficiencies for the swept ramp injector configuration were
close to the values for the perpendicular sonic injection mixing model
and were independent of the freestream total temperature.
Expansion ramp injectors were designed to reduce the pressure loss
associated with the oblique shock produced by compression ramps
Fig. 31. Flowfield of an expansion ramp injector [286]. [294]. In the flowfield of an expansion ramp, as shown in Fig. 31, the
flow on the top of the ramp expands toward the ramp edges starting from
fraction distribution after a swept ramp injector. An air jet (Mj = 1.7) the local Mach angle location. The flow on both sides of the ramp is
from the ramp was injected into supersonic airflows (M∞ = 2). The directed vertically toward the expansion surfaces, and a small pressure
injectant plume was distorted and lifted away from the injector wall by gradient develops near the ramp edge. The flow moving down at the side
vortices generated by the ramp injector. In the far field of the ramp in­ of the ramp results in a vorticity generation mechanism known as the
jection, large-scale coherent streamwise vortices dominated the mixing cross-stream shear [286]. Two counter-rotating streamwise vortices are
process over approximately ten ramp heights downstream of the formed at a short distance downstream of the injector base. The
injector, then flow turbulence started to take over the mixing. Honig expansion surface allows effective fluid entrainment into the fuel jet by
et al. [293] visualized the flame structures of a swept ramp injector in a increasing its contact surface area and the instability of the fuel-air shear
hydrogen-fueled supersonic combustor. A hydrogen jet (Mj = 1.6) was layer [295]. A recompression shock develops where the flow is turned

28
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 32. Contoured wall injectors [296,297].

Fig. 33. Unswept and swept hypermixing nozzle configurations [300].

Fig. 34. Wall-mounted alternating wedge ramp injectors [307].

parallel to the freestream at the base of the ramp. Due to the expansive Contoured wall injectors were designed to increase the mixing pro­
flowfield, the injected fuel can only be ignited by an external igniter and cess by shock-induced baroclinic effects. Marble et al. [296] proposed
does not burn until a certain distance downstream of the injectors [294]. the concept of a contoured wall nozzle for rapid mixing and combustion

29
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

of hydrogen and air in the supersonic flow, as shown in Fig. 32. The introduced. Nevertheless, flow separation upstream of the expansion
injector consists of alternate compression ramps and expansion troughs. ramp can result in the deterioration of hypermixer mixing performance.
At the end of each ramp, a fuel nozzle is used to discharge fuel parallel to Kubo et al. [306,307] applied the wall-mounted alternating wedge ramp
the supersonic flow. The flow in the channel between the ramps is injector to improve the performance of the scramjet combustor in a
turned parallel to the freestream in the plane of injection. An oblique direct-connect combustion facility (M∞ = 2.7). It was observed that the
shock forms between the ramps at the base of the expansion troughs. combustor performance was substantially decreased at high fuel
This shockwave intersects the density gradient existing between the equivalence ratio conditions due to flame attachment to the injector.
injected fuel and incoming air generating streamwise vorticity via bar­ When the flame was stabilized downstream of the injector, the flow on
oclinic torque. The vorticity coalesces into a counter-rotating vortex pair the surface of the injector expansion ramp was not separated, which
and migrates the fuel away from the wall to enhance both mixing and jet facilitated the streamwise vortices generation in the injector wake re­
penetration [297–299]. Nevertheless, Waitz et al. [299] found that the gion. When the flame was attached to the injector at high fueling rates,
primary source of axial vorticity in the flowfield was not baroclinic the injector wedges and expansion ramp were covered by a flow sepa­
torque. Numerical analysis showed that the shock-induced streamwise ration region, which caused a significant decrease in combustor
vorticity was only 30% of the total streamwise vorticity developed by performance.
the injection geometry. The majority of the streamwise vorticity was Intrusive injectors (e.g., strut and pylon) have been used for in-
generated by the variation in aerodynamic loading along the surface of stream fueling and enhanced mixing in supersonic flows. A strut spans
the injector ramps, which was the same as in the lobed mixture flows. the entire flowpath, i.e., the full height or width of the cross section in
Davis and Hingst [300] combined the features of compression and combustors, to deliver the fuel into the core flow region. A pylon has its
expansion ramps and contoured wall injectors as shown in Fig. 33. Flow height deep into the fuel core but does not span the entire flowpath.
passing through the compression-expansion ramp injectors is initially Simple strut and pylon injectors are characterized as fuel placement
compressed by the planar oblique shock generated by the compression devices rather than fuel mixing devices [246]. These injectors can
ramp. The top surface of the ramp injectors remains co-planar with the potentially be effective for fuel delivery in a large-scale combustor
compression ramp. The flow parallel to the compression wall splits where the ability for ramps or flush-wall injectors to access the core flow
where the ramp surface changes. A pair of counter-rotating vortices is is limited [246]. The integration of streamwise vortex generators on
generated when the high pressure air on the upper surface of the ramp strut and pylon injector configurations has been extensively studied to
spills over the ramp side edges into the low pressure region created by enhance the fuel dispersion and mixing. The streamwise vortices pro­
flow expansion, which is similar to the simple compression ramp duced by in-stream injectors can enhance the supersonic mixing due to
injector. At the end of the expansion ramp, another set of oblique shocks the following reasons: (1) the large-scale streamwise vortices are easily
is produced when the flow is turned parallel to the freestream. Experi­ generated and free from compressibility effects in the azimuthal motion,
mental results [300] demonstrated that the swept ramp model produced (2) the flow circulation in azimuthal flow direction and spatial posi­
stronger and larger vortex pairs than those produced by the unswept tioning can be easily controlled, and (3) vortex control in large scales
ramp model. The interaction between the vortex pairs and the boundary and subsequent interaction between the streamwise vortices in small
layer caused the center jet of the swept ramp model to mix faster than scales can control the turbulent flow structures in combustors [308].
the outboard jets. However, the swept ramp model had a higher total In-stream strut/pylon injectors are candidates for fueling devices at the
pressure loss. low corridor of hypersonic flights to enhance supersonic mixing by
Itoh et al. [301–305] proposed a compression-expansion type introducing both fuel and streamwise vortices directly into the core of
hypermixer, a wall-mounted alternating wedge ramp injector, for the supersonic flows.
M12-02 scramjet developed by JAXA. As shown in Fig. 34, this hyper­ Fig. 35 shows the alternating-wedge struts used to explore the effects
mixer has a 2D compression ramp at the upstream side and a following of streamwise vortex arrangements on mixing characteristics [308].
section to generate streamwise vortices. On the downstream side, Struts are characterized by their alternating upward and downward
wedges are extended from the upstream compression ramp and arranged expansion ramps in the spanwise direction of the strut’s main body. The
alternating with expansion ramps in the spanwise direction. In this strut generates counter-rotating or co-rotating streamwise vortices in a
configuration, a cavity is formed under the compression wedge by the spanwise row configuration at the boundary of neighboring expansion
composition of alternating ramp wedge. The cavity can enhance the ramps. Swept angle applied to the ramp can increase the circulation of
generation of small-scale turbulence eddies into large-scale streamwise the streamwise vortex compared to the parallel ramp. Gaseous fuels are
vortices after the intrusive wedges. The hypermixer performance can be injected parallel to the freestream direction directly into each stream­
increased by swept ramps when streamwise vortices are not sufficiently wise vortex core through series of sonic orifices located at the boundary

Fig. 35. Alternating wedge strut injectors [308].

30
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 36. Pylon injector configurations [315].

geometry on the others. The mating of the two pieces creates thin
rectangular slots for sonic fuel injection. The basic pylon embodies two
fuel-injection strategies: (1) maximizing the fuel/air interface area with
a rectangular slot and (2) injecting fuel upstream of the pylon aft area to
allow fuel and air mixing before leaving the pylon surface.
Doster et al. [315] characterized the mixing performance and total
pressure loss of pylon injectors in Mach 2.0 airstreams using Raman
spectroscopy and NO-PLIF. Fig. 37 compares the instantaneous and
ensemble-averaged NO-PLIF images of fuel plumes in non-reacting
flows. The fuel plume from the basic pylon was highly unsteady and
exhibited side-to-side movement about the centerline of the plume due
to vortex shedding, whereas the wake-region flow of the two hypermixer
pylons was much steadier than that of the basic pylon. The increase in
streamwise vortical motion resulted in better mixing for the ramp and
alternating-wedge pylons, bringing a larger area of fuel/air mixture to
flammable conditions over a shorter distance. The mixing enhancement
of the hypermixer pylons moved the combustible mixture region closer
to the base plane of the pylon compared with the basic pylon. There was
a measurable increase in total pressure loss accompanying the increased
mixing efficiency in the case of the alternating-wedge pylon compared
with the basic pylon, while the ramp and basic pylons had similar total
pressure losses. These studies were focused on the comparison of the
performance of the pylons in terms of the total pressure losses and
mixing levels between the three pylon injectors, but no attempt was
made to understand how the interaction among the vertical structures
modified the combustion zone in the jet plume.
Vergine et al. [316,317] explored the effects of the imposed inter­
Fig. 37. NO-PLIF images downstream of pylon-type hypermixers [315]. action and dynamics of the supersonic streamwise vortices on the
reacting plume and its evolution. The ignition and combustion charac­
of the neighboring expansion ramps. A series of experimental [308–310] teristics of the hydrogen plume from pylon injectors were characterized
and numerical studies [243,309,311,312] were conducted to charac­ at Mach 2.4 in a high-enthalpy expansion tube facility. The pylon in­
terize the supersonic mixing enhancement and flame stabilization by jectors have symmetric and asymmetric expansion ramps added to the
alternating-wedge strut injectors in Mach 2.5 airstreams. In summary, it basic fin body that enhances the vortex dynamics. The fuel was delivered
was shown that the jet plume issuing from the struts with axial vortex by a thin slit nozzle located at the centerline of the trailing edge, which
generation was shaped by streamwise coherent structures in early spanned the entire height of the pylons. Fig. 38 shows the evolution of
development and large-scale counter-rotating vortices in flow regions the reacting plume of symmetric and asymmetric configurations at
downstream. The streamwise vortices modify the combustion stability different stations downstream of the injection point. In the symmetric
and ignition characters of the jet plume. The streamwise vortex circu­ case, the vortical structures evolving from the ramps interacted,
lation was affected by the ramp angle, spanwise spacing, sweep feature, generating two distinct quadruplets of vortices as they moved down­
and vortex interactions across the symmetry planes. The optimization of stream. The persistence of the lobes and the repetition of the same flow
injector design was lacking even though the major modification was features in each section indicate that the coherent structures have an
introduced in the jet plume by streamwise vortices. important effect on the dynamics driving the plume evolution. In the
Fig. 36 presents pylon-type hypermixers for in-stream enhanced asymmetric case, the plume showed a characteristic shape, i.e., a wide
mixing in scramjet combustors, which includes the basic pylon, the ramp lobe and two smaller areas where hydroxyl radicals agglomerated in the
pylon, and the alternating-wedge pylon [313–315]. The forward area middle and at the extremities, respectively. The vortices entrained fuel
contains a plenum common to all injectors. The aft area forms a and air in the right proportions to ensure a homogeneous combustion
constant-angled compression ramp on the basic pylon and a hypermixer and enhanced the plume’s spreading in their area of influence. In
conjunction with numerical simulations, it was found that the

31
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 38. Instantaneous flame distributions of symmetric (left) and asymmetric (right) cases at cross planes to the fuel port a) 4.3 cm, b) 7.6 cm, c) 10.7 cm [317].

coalescence of vortical structures of the same rotating direction formed While rapid mixing is necessary for producing effective heat release
in the center of the plume, and those convecting downstream at the using pylon injectors, the associated cost of mixing enhancement on
upper and lower edges of the plume shaped the evolution of the com­ stagnation pressure loss should be alleviated. There exists a tradeoff
bustion zone. between efficient mixing and maintaining stagnation pressure, which is

32
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

of fin-guided fuel injection using a comparison between transverse fuel


injectors with and without a fin upstream.
Aguilera et al. [318,319] investigated fin-guided fuel injection and
mixing in a Mach 2.1 airflow in a blow-down supersonic tunnel. Helium
gas was injected sonically into the supersonic crossflow to simulate
gaseous hydrogen injection. The fin pylon has a triangular blade just
upstream of the orifice injector on the lower wall such that the injector
orifice is just downstream of the fin’s trailing edge, allowing fuel in­
jection into the fin’s wake. Fig. 40 presents Mie-scattering images of
gaseous fuel dispersion at multiple cross-sectional locations downstream
of the baseline injection (without a fin) and the fin-guided injection
[318]. Due to the steepening of the triangular cross-section of the fin, a
pair of counter-rotating vortices was generated behind the fin, which
enhances the fine-scale mixing. Comparing the corresponding images of
the baseline and the fin-guided cases, much of the fuel in the fin-guided
injection case was delivered to the supersonic core rather than to the
boundary layer as in the baseline case. The cross-sectional images after
the fin-guided injection showed jet curling phenomenon due to the ef­
fect of the fin-generated axial vortices on the helium jet. Wall pressure
measurements showed that the injection shock strength associated with
fin-guided fuel injection in the non-reacting case was significantly
reduced. The wall pressure reduction reached nearly 33%–47%
depending on the injection pressure. The benefits of fin guided fuel in­
jection decreased with increasing fuel flow rates.
Liu et al. [320] compared the flow structures in pylon-aided injection
Fig. 39. Comparison of jet penetration behind a jet-induced shock and a fin- and normal injection cases of a sonic jet into a Mach 2.95 supersonic
initiated shock [318]. crossflow. Oil flow visualization was used to study the detailed in­
teractions between the crossflow and the jet with the pylon installed
closely associated with propulsion system efficiency [318]. Slender upstream of the orifice as shown in Fig. 41. Upstream of the orifice, the
fin-type pylons are designed as a fairing mounted just upstream of the crossflow is separated by the pylon and encounters the 3D bow shock
fuel injector with the maximum base width less than or equal to the induced by the jet obstruction. A separation shock is generated by the
injector diameter. The fin creates a low-pressure region to increase the upstream horseshoe and intersects with the detached shock appearing at
penetration depth of a transverse jet downstream. The fin would not add the leading edge of the pylon. Downstream of the jet, the shockwave
additional flow blockage to that naturally incurred by the transverse fuel surface originates from the collision of the supersonic flow surrounding
jet. The fin-pylon emphasizes the mixing enhancement as well as the the barrel of the jet. The collision shock interacts with the reflected
reduction of drag and total pressure loss. Fig. 39 illustrates the concept shock produced by the barrel shock. The induced recirculation zone at

Fig. 40. Cross sectional views (Mie-scattering images) of gaseous fuel dispersion comparing the baseline and the fin-guided cases [318].

33
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 41. Oil flow images of jet injectino in the supersonic crossflow at J = 20.6. The upper part is for the normal injection case and the lower part is for pylon-aided
one [320].

the jet leeward edge extends in the spanwise direction after the incident interaction position. Pylon-aided injection cases had a longer distance
collision shock. Flow structures downstream of the jet between the from the orifice to the shock interaction position and a larger jet leeward
normal injection and the pylon-aided injection cases had different sizes separation zone.
of the V-shape separation zone (the ending points are labeled by red Another type of pylon-type injector, the delta wing shaped vortex
dots) and the shock (the collision shock and the reflected shock) generator [321], has been considered to improve the mixing between
the incoming supersonic air and the fuel injected downstream. The delta
wing height and the jet-to-crossflow pressure ratio are parameters
dominating the mixing efficiency and fuel penetration depth. Li et al.
[321] found that the mixing efficiency decreased with the increase of the
jet-to-crossflow pressure ratio irrespective of the height of the delta
wing, and there was an optimum height of the delta wing for each
jet-to-crossflow pressure ratio to maximize the rapid fuel–air mixing.
There are practical difficulties with operating ramp-, strut-, and
pylon-type injectors in extremely high enthalpy flows, though the mix­
ing characteristics of these types’ injection may be attractive [322]. In a
scramjet combustor, the harsh flow conditions and shock waves may
pose a severe threat to the physical integrity of an obtrusive type in­
jection system. For instance, the streamwise vortex generation from a
ramp injector depends primarily on pristine geometry and sharp corners,
which cannot be sustained over long periods of time in the environment
of a high-enthalpy scramjet combustor. All intrusive fuel injectors create
hot spots on the combustor walls with temperatures exceeding the
thermal limits of most practical materials. The cooling systems needed
for intrusive injectors increase the complexity of the fuel injection sys­
tem and the overall structure efficiency. Besides, the presence of flow
blockage from intrusive injectors adds to the aerodynamic drag, the total
pressure loss, and thereby pose a threat for thrust degradation, unac­
ceptable in operating the engine at high Mach numbers [289]. Alter­
natively, an array of wall-flushed port injectors can be arranged in such a
way as to simulate the aerodynamic pattern of the intrusive injection
scheme accomplishing this mixing augmentation without the
complexity, losses and physical integrity problems associated with the
various kinds of struts and ramps that have been suggested in the past
[322]. Experiments and computational studies have demonstrated that
the aerodynamic ramp injection system has better mixing characteristics
than a single transverse injection system and would be a realistic method
Fig. 42. Schematic views of (a) mean flowfield, and (b) flame distribution in a
for fuel-air mixing in scramjet combustors.
dual transverse injection system [323,324]. The flowfield will necessarily have multiple or staged-injection

34
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

configurations when using a plasma torch or pilot torch as an igniter or


using a pilot injection of hydrogen for combustion stabilization of hy­
drocarbon fuels. An efficient multiple injection system includes opti­
mization of (1) each injector position, (2) distribution of mass flow rate
and momentum flux, and (3) the combination of injection angles. It is
challenging to analyze the mixing and combustion characteristics of a
system with multiple transverse injection. These systems can be
simplified and represented by a basic model of a dual transverse injec­
tion system, in which the injectors are located in a line.
Fig. 42 shows the mixing and combustion characteristics of a dual/
staged transverse injection system [323,324]. Staged injection generates
many shockwaves including 3D bow shocks formed ahead of the pri­
mary (front) jet and ahead of the secondary (rear) jet, a separation shock
generated by the interaction between the front bow shock and boundary
layer, and Mach discs. The front jet blocks the supersonic air flow and
creates a low pressure region to increase the penetration depth of the
rear jet. The Mach disc of the secondary jet is located at a higher position
from the wall and is larger than the Mach disc of the primary jet. The Fig. 43. Concept of an aerodynamic ramp injector [333].
evolution of vortex pairs formed along the jet flows, horse-shoe vortices,
separation bubble and recirculating wake flows involves 3D tilting and
flow were revealed using two-point correlation maps from the acetone
folding, which accelerate the fuel-air mixing in a supersonic flowfield
PLIF images. The secondary injection influenced the large-scale struc­
and thus enhances global mixing characteristics. The secondary jet en­
ture in size and inclination angle. Takahashi et al. [327] further inves­
hances the fuel/air mixing by promoting the development of large-scale
tigated the mixing enhancement by pulsed secondary injection at
structures, which are formed by the interaction between the primary jet
identical flow conditions. The pulsed jet influences the primary jet
and the main air flow. Aerodynamic blockage effects have significant
through the change of the separation region between them. The primary
influence on combustion characteristics. The heat addition due to the
jet was completely lifted up in the pulsed-injection case due to the
burning of the primary injection might lead to a significant change in the
pressure rise caused by the pulsation. The detailed mechanism of the
flowfield between injectors and thus lead to a change of blockage effect
interactions between the counter-rotating vortex pair and the near-wall
with respect to the nonreactive case. The most important parameters
separation region around the secondary jet is not clear.
affecting the blockage effects and the evolution process of streamwise
Performance of staged fuel injectors has been comprehensively
vortices are the distance between injectors and the jet-to-crossflow
explored using numerical simulation in terms of the overall mixing
momentum flux ratios. There exists an optimal distance between in­
characteristics, streamwise vortices, mixing rates, penetrations at
jectors for the mixing, and the optimal distance would change according
different jet momentums, and optimal distances for mixing and com­
to the jet-to-crossflow momentum flux ratios. Additionally, the pre­
bustion stabilization. Lee [323,324] characterized the mixing and
heating of the primary jet can augment flow expansion and improve the
combustion of a dual sonic hydrogen injection into a Mach 2 airstream.
ignition and burning of the secondary injection flow.
The effects of the jet-to-crossflow momentum flux ratio and the distance
Hollo et al. [325,326] conducted a 3D survey of staged transverse
between injectors were investigated systematically. It was found that the
injection into a Mach 2 freestream. The planar laser-induced iodine
rear injection flow was strongly influenced by the blockage effect from
fluorescence (LIIF) was used to provide quantitative distributions of the
the momentum flux of the front injection flow, and thus had higher
injectant mole fraction and velocity in a nonreacting model combustor.
expansion and penetration. The burning process of the rear injection
The measurements revealed the effect of streamwise vortices on the
benefited from both the blockage and the preheating effects due to
mixing of the injectant in the near field of the injectors, as well as the
chemical reactions of the front injection flow. Landsberg et al. [329]
rapid mixing generated by staging two fuel injectors. Vortices were
explored the effects of cascading hydrogen injectors at combustor con­
generated as the freestream spilled around the underexpanded jet cores.
ditions corresponding to Mach 6–12 flight. The downstream injector
They promoted rapid mixing of the injectant by driving freestream air
benefited jointly from the shielding effect induced by the smaller up­
into the center of the jet plume. The staged injectors produced a
stream injector and its increased diameter. Unique optimal injector
significantly higher initial rate of injectant mixing than a single injector
spacings existed for each freestream condition, however, no universal
due to the interaction between the jet cores. In the far field of the staged
optimum spacing was determined. Other numerical studies focused on
injector, the rate of the dominant small-scale turbulent mixing was in­
the mixing enhancement and drag reduction through multiport injector
dependent of injection geometry, freestream Mach number, and the
arrays [330–332].
injectant molecular weight. The transition of the dominant mixing
Aerodynamic ramp injectors were proposed as non-obtrusive axial
mechanism from vortex-driven mixing in the near field to small-scale
vorticity generators to simulate physical ramp injection in harsh
turbulent diffusion in the far field occurred in the region about 10 jet
combustor environments [322]. Fig. 43 shows an aerodynamic ramp
diameters downstream of the injectors.
injector, which consists of nine flush port injectors arranged in a 3 × 3
Takahashi et al. [327,328] examined the behavior of large-scale
matrix symmetric about the streamwise axis. The injector array is
structures and their role in the mixing of the staged injection configu­
designed to produce vortical interactions to enhance mixing in a su­
ration. The primary jet was helium and the secondary jet was air, and
personic crossflow from a flush wall injector. The most upstream row of
they were transversely injected into an unheated Mach 2.0 crossflow at
injectors is at a low transverse angle, and the transverse angle of the port
sonic conditions. Acetone PLIF was used to quantify the molar concen­
injectors increases along the flowpath to create the ramp-like aero­
tration of the acetone seeded in only the primary jet of the staged in­
dynamic feature. The outside port injectors in the second and third rows
jection, which had sufficient capability to resolve the large-scale
are displaced and yawed inward stepwise to create the aerodynamic
coherent structures. The single-time two-point spatial correlations of the
sweeping further. The mixing characteristics of the physical ramp in­
injectant concentration based on the concentration fluctuation were
jection are dominated by the counter-rotating vortices generated by the
used to characterize the size, shape, and orientation of the large-scale
obtrusive ramp. The fuel-air mixing produced by the aerodynamic ramp
structure and its motion. The large-scale structures on the upper jet
is dominated by the multiplicative fuel-vortex interactions confined to
boundary region and its dominant role in stirring the jet and the main

35
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 44. Surface oil-flow visualizations of an aerodynamic ramp in a Mach 2.4 crossflow [337,338].

the injector vicinity [334]. Hence, the mixing characteristics depend on jet-to-freestream momentum flux ratio of 1.1–3.3. The aerodynamic
the injector array spacings, injector yaw angles, injector transverse an­ ramp injector had two rows of two round holes, spaced 4.0 equivalent jet
gles, and dynamic pressure ratios. Parametric numerical analysis [333, diameters apart in the streamwise direction with a cross-stream spacing
335,336] indicated that neither inward displacement nor yaw of the of 2.0 equivalent jet diameters between the holes. A single low-angled
outer-column injectors relative to those upstream was necessary for the circular injector was used as the reference injector having the same
production of strong streamwise vortices by the aerodynamic ramps. total injector area. Fig. 44 shows the surface oil-flow patterns created
The inward displacement and yaw enhance vertical penetration of fuel around the two injectors [337–339]. As the jets expand inward, they
but reduce the overall fuel-air mixing characterized in planes down­ reduce the width of the overall shock and confine the blockage to a
stream normal to the flow direction. Meanwhile, increasing the trans­ narrower cross-stream space with a higher overall plume penetration.
verse angle of injectors at the symmetric centerline improves both Three separation zones were created in front, in between, and behind
penetration depth and mixing efficiency but strengthens jet-induced
shocks and thereby increases total pressure losses. Increasing the dy­
namic pressure ratio produces a similar effect to increasing the injection
angle.
Mixing of aerodynamic ramp injectors has been investigated in
nonreacting flows. Cox et al. [322] originally presented the novel
aerodynamic injector array and characterized the mixing effectiveness
based on maximum fuel mass fraction, plume area, mixing efficiency,
and spatial mixedness. Sonic helium jets were injected into a Mach 3
airstream at underexpanded pressure conditions to simulate hydrogen
fuel injection in scramjets. The mixing rate achieved by the aerodynamic
ramp injector was found to be superior to that reported for the physical
swept ramp injectors at 2.5 times the pressure ratio of jet to crossflow at
the underexpanded pressure condition. Fuller et al. [289,334] experi­
mentally compared the supersonic mixing performance of an aero­
dynamic ramp injector with that of an equivalent physical ramp injector
in terms of jet penetration, mixing characteristics, and total pressure loss
in a Mach 2 airstream. It was shown that the aerodynamic ramp pro­
duced superior mixing in the near field and inferior mixing in the far
field, while the aerodynamic ramp mixing approached that of the
physical ramp in the far field with increased jet momentum. Increasing
the jet momentum reduced the effect of vortices produced by the
physical ramp, whereas it increased the strength of interactions pro­
duced by the aerodynamic ramp. Hence, the overall mixing performance
decreased for the physical ramp, while increasing for the aerodynamic
ramp with increasing jet momentum. The physical ramp suffered a more
severe overall pressure loss than the aerodynamic ramp.
Jacobsen et al. [337–339] explored the flow structures of a multiport
aerodynamic ramp injector array in a Mach 2.4 airflow with a
Fig. 45. Normalized growth rate of natural and excited shear layers [358].

36
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

the injector holes. The buildup of oil in the separation zones in front of energy deposition [354], acoustic driver, piezoelectric actuator, suc­
the aerodynamic ramp injector holes leaked through between the first tion by counterflow, and jet deflection (e.g., flip-flop nozzle, flapping
row of injector holes at low momentum flux ratios, and the surface flow jets) [355–357]. Nevertheless, passive/active methods specifically
was stopped at higher momentum flux ratios. Initially the separation devised to excite the shear layer instability of jets may not be applicable
zone between the two rows of injector holes helped the jet plume in scramjet engines due to the high amplitude needed to modulate the
penetrate farther by reducing the influence of the freestream momentum high-enthalpy and high-Mach number flow. There are limited studies of
on that of the jets. The separation zone behind the rear injector holes such types of shear layer excitation methods used in scramjet engines.
increased in size with momentum flux ratio. The relatively large area of A cavity is considered as one of the best acoustic drivers to generate
mixed fuel and air between the aero-ramp plume cores, coupled with the acoustic waves strong enough to disturb the stable supersonic flow
low-velocity region behind the rear separation zone of the aero-ramp, [359]. Flow-induced cavity resonance which can excite shear layer
could provide a suitable region for flame initiation and rapid instability has been observed widely in cavity flameholding devices in­
spreading into the jet plume. The plume of the aerodynamic ramp had a side scramjet combustors. This effect can generate large-scale lateral
larger area than the single-hole injector due to lateral spreading and vortices to increase the rate of entrainment and dominate turbulent
faster mixing, which led to faster decay of the maximum mixing con­ compressible mixing and combustion. The spreading rate of the excited
centration for the aerodynamic ramp. The aerodynamic ramp plume had shear layers is substantially higher than the natural growth rate of the
a higher initial mixing core penetration height but was overtaken by the turbulent compressible shear layer [358], as shown in Fig. 45.
single-hole injector at the farthest downstream location. Cavity-induced mixing enhancement can be critical in combustion sta­
bilization in terms of the mixing efficiency and extending the combus­
2.2.4. Excitation of shear layer instabilities tion stability limit over the entire combustor.
Excitation of shear layer instabilities have been widely used to Yu et al. [340,341,360–362] investigated the excitation of shear
enhance flow mixing in combustion devices. Passive and active excita­ layer instability of an expanded Mach 2.0 circular jet using open rect­
tion techniques have been devised to excite the shear layer instability of angular and semi-annular cavities. The supersonic jet was acoustically
supersonic jets to enhance the air entrainment into the jet core and forced by rectangular and semi-annular cavities mounted adjacent to the
hence the combustion efficiency. The growth rate of the shear layer can jet exit plane with the open side facing the shear layer. Fig. 46 shows the
be increased by passive excitation using cavities [340,341], acoustic large-scale coherent structures shedding downstream of the cavity. The
feedback [342], nozzle attachments (e.g., collars, ramps) [343], inter­ growth rate of the shear layer was substantially increased by the
action between jet shear layer instability and unstable wakes originating large-scale highly coherent structures that were produced when the
from the nozzle rim [344,345], and interaction between multiple su­ cavities were tuned to certain frequencies, which depended on the cavity
personic jets at resonant conditions [346–352]. Active excitation length and depth. The observed forcing frequency could be explained by
methods include glow discharge [353], electric arc and pulsed-laser either a convective-acoustic feedback mechanism or normal mode
resonance of the cavity. The cavity-actuated forcing increased the
spreading rate of the initial shear layers by a factor of three for a cold
Mach 2.0 air jet with a convective Mach number (Mc) of 0.85, whereas a
50% increase in the spreading rate was observed for high-temperature
Mach 2.0 jets with Mc of 1.4. The increased growth rate of the shear
layer resulted in a 20%–30% reduction in the afterburning flame length
in the reacting supersonic jets with ethylene-oxygen afterburning.
Sato et al. [359] reported an experiment of mixing control in a Mach
1.78 airstream through a wall-mounted cavity. The acoustic wave from a
rectangular cavity impinged on the initial condition of the supersonic
mixing layer at the 2D nozzle exit to affect its growth. The experiment
results of shear layer growth from pitot pressure profiles confirmed the
possibility of enhancing the 2D structure evolution by acoustic distur­
bance. Nenmeni and Yu [363] characterized cavity-induced mixing
enhancement over a wide range of stagnation pressure conditions that
may be encountered during a supersonic flight. It was found that the
frequency and amplitude of cavity-induced oscillations were critical
parameters in producing large-amplitude coherent fluctuations for
practical mixing enhancement.
Though a wall-mounted cavity can enhance the mixing by acoustic
excitation, it is usually used as a flameholding device in scramjet com­
bustors. The flameholding characters and mixing enhancement are
coupled in high-speed reacting flows and could be changed by cavity
geometry. Yu et al. [358] evaluated the effects of acoustically open
cavities on an inclined sonic jet in a Mach 2 airstream. The fuel was
injected into the supersonic freestream at a 45◦ angle, and cavities
having various sizes and aspect ratios were placed just downstream of
the fuel injection point. Fig. 47 shows spark schlieren images taken at
various locations. Those images were combined to construct an
ensemble view for each configuration. Strong acoustic oscillations were
detected in the no. 4 cavity, while no organized pressure oscillations
were detected inside the no. 2 or no. 5 cavities. The cavities with a short
aspect ratio provided good flameholding, whereas those with a rela­
Fig. 46. Schematic of the cavity-actuated supersonic jet forcing (top), and tively long aspect ratio shortened the flame length substantially via
planar Mie-scattering images of a Mach 2.0 jet, natural (middle) and forced acoustic excitation. With the use of unstable cavities, the downstream
with a rectangular cavity (bottom) [341]. shock structure became more unsteady and irregular. The combustor

37
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 47. Spark schlieren images of fuel injection in various cavity configurations [358].

wall pressure and the exit recovery temperature measurements showed shear layer instability excitation and mixing might not be appropriate
that the volumetric heat release was substantially increased by cavities for stable scramjet operation and ram-to-scram mode transition. The
compared to the baseline case without a cavity, indicating better mixing. unsteady nature of the free shear layer is suppressed by the angled back
Mixing enhancement by excitation of shear layer instabilities has wall design in cavities, which eliminates the generation of travelling
limited application in scramjet engines although various excitation shocks inside the cavity due to free shear layer impingement. The
techniques have been devised and widely used in subsonic and super­ detailed mechanism of cavity flameholding and relevant physical issues
sonic combustion jets. Meanwhile, the resonant cavity optimized for are further discussed next in the section on cavity flameholding.

2.3. Physical flame holding

Combustion stabilization has been the fundamental challenge in the


design of scramjet engines including fuel injection, ignition, and
flameholding [87]. At the low range of the hypersonic flight envelope,
between Mach 5–8, the relatively low stagnation temperature cannot
Fig. 48. Flameholders in scramajet combustors: step, cavity, and strut [106]. auto ignite the fuel, and the associated long ignition delay may become a

38
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

critical issue for combustion stabilization [96]. Cavity flameholders consists of both broadband small amplitude pressure fluctuations typical
have been proposed as integrated fuel injection and flameholding de­ of turbulent shear layers as well as intense discrete resonances driven by
vices for supersonic flameholding and combustion performance vortex shedding from shear layer and acoustic feedback. The frequency,
improvement in scramjet engines due to their minimal system amplitude, and harmonic properties of cavity resonance depend on the
complexity and performance deterioration (i.e., total pressure loss). Wall cavity geometry and external flow conditions.
cavities, steps, and strut-type flameholder in scramjet combustors can Plumblee et al. [395] systematically investigated the acoustic
increase flow residence time and enhance the fuel-air mixing rate by response of cavities in both subsonic and supersonic airflows. Both the
creating recirculation regions and pools of hot radicals that reduce the random frequency buffet response and the discrete-frequency resonant
induction time of chemical reactions. Recirculation regions created by response were observed. The buffet response of the cavity was charac­
those flameholders can also serve as self-sustaining ignition sources terized by a random spectrum which reaches its maximum value at the
[364]. Research activities on flameholding mechanisms of supersonic lower limiting frequency of the analyses. It was hypothesized to result
combustion have been performed at WPAFB (U.S. AFRL), University of from the unstable nature of the separated flow, which tends to permit an
Illinois (U.S.), University of Michigan (U.S.), DLR (Germany), JAXA intermittent direct impingement of flow on the rear face of the cavity.
(Japan), China Aerodynamics Research and Development Center, and Theoretical analysis assumed that the cavity may pick up certain fre­
National University of Defense Technology (China). In this section, quencies to amplify from the broad band noise source produced by the
comprehensive understandings of flameholding mechanisms are sum­ turbulent shear layer spanning the boundary of the cavity. Since cavity
marized from literature. acoustic response allows only those modes which meet the boundary
conditions to form a standing wave, the results of the excitation of the
2.3.1. Physical flameholders normal acoustic modes in the cavity are the discrete frequency oscilla­
In scramjet engines, commonly used physical flameholding methods tions. It was suggested that the oscillatory frequencies of a cavity flow
can be categorized into non-intrusive and intrusive methods. In non- were identical to those which correspond to the maximum acoustic
intrusive flameholders, a wall-mounted cavity [96,97,364] or a frequency response of the cavity. The characteristic frequency equation
rearward-facing step [365–368] creates a low-speed recirculation zone is
outside of the supersonic core flow. As the flame is close to the wall, [( )2 ( )2 ( )2 ]
extra attention should be paid to the wall’s thermal protection. Intrusive c2 nx ny nz
fN2 = + + , (5)
flameholders, i.e., strut [369–372] and bluff body [259,373–377], are 4 Lx Ly Lz
placed into the supersonic core flow to promote fuel-air mixing and
where L is the cavity dimension, n is the mode number, c is the speed of
create a subsonic region within the mainstream. The stabilized flames
sound, f is the resonant frequency, and x, y, z are the Cartesian co­
and combustor walls are separated by the high-speed flow. Efficient
ordinates in streamwise, spanwise, and cavity depth directions. Trans­
cooling methods (ablation and regeneration) should be applied to pro­
verse oscillation frequencies of the cavity were predicted accurately by
tect the integrity of instream struts. Due to the common nature of
the analytical model for cavities with L/D (length/depth) ≤ 1 only at all
flameholding by creating a low-speed recirculation regions in supersonic
Mach numbers. This mode has not been justified for the longitudinal
freestreams, the three different types of physical flameholders, (i.e., step,
oscillatory modes of the cavity flow. Nevertheless, this analysis has met
wall cavity, strut) share basic features: a shear layer, a recirculation
obvious difficulties when the oncoming boundary layer of a cavity flow
zone, and an outer freestream [106], as shown in Fig. 48. In high-speed
is laminar [273,394,396]. Experiments revealed that laminar flow pro­
combustion devices, fuel can be directly injected into the recirculation
duced stronger oscillation even though the broad band excitation as
regions or can be entrained from the jet injection upstream of the
required by the Plumblee et al. model was absent.
flameholder. The combustion is essentially non-premixed in the direct
Explanations of the longitudinal cavity oscillation are based on the
fueling configuration or partially premixed in the upstream fueling
interaction between the shear layer and the cavity edges. The shear layer
configuration.
impingement on the rear wall increases the cavity pressure and creates
These three different flameholding devices can be integrated to form
an acoustic wave that propagates upstream at the local speed of sound
combined flameholding schemes, which provide synergistic methods for
and impacts the front wall. The shear layer deflects upward and
overall combustor performance improvement in scramjet engines. The
downward due to the acoustic perturbation from the cavity. The shear
non-intrusive flameholding devices have limited fuel delivery capability
layer oscillation is self-sustained when the shear layer deflection prop­
while the intrusive flameholding devices penetrate into the supersonic
agates downstream and results in another impingement event on the
core flow and significantly increase the fuel penetration depth. Due to
rear wall of the cavity, providing a feedback for continuous shear layer
the variance of the fuel penetration capability between the cavity/step
deflection. The unsteady motion of the shear layer results in the periodic
and strut/bluff body, the combined physical flameholder can achieve
mass addition and removal at the cavity trailing edge. There are two
efficient fuel mixing, promote fuel ignition, and extend the flame sta­
primary models used to explain the acoustic perturbation leading to
bility limit. For instance, the cavity-bluff body or cavity-pylon combi­
shear layer instabilities in the longitudinal cavity oscillation process.
nation flameholding scheme can broaden the operation range of a
The first model assumes that the acoustic wave reflected by the front
simple cavity flameholder by instream strut-wake generation [318,319,
wall induces small vortices at the cavity leading edge, which grow as
378–386]. A tandem-type strut-cavity flameholding scheme extends the
they are convected downstream and result in the shear layer deflection
cavity influence into the core flow [369,387]. Two parallel ramp-shaped
upward and downward. The second model presumes that the acoustic
struts significantly improve the combustion and mixing efficiency of the
wave reflection from the front wall is the cause of the shear layer
combustor when compared with single strut configuration [388]. Staged
deflection and the impingement event on the rear wall.
struts and multiple struts are also observed to reduce the lean blowout
Rossiter [397] investigated the flow over rectangular cavities at
limit of supersonic combustion [389–393]. Nevertheless, characters of
subsonic and transonic speeds. It was found that the random component
simple physical flameholders, mostly the cavity flameholder, are
predominates in the shallower cavities (L/D > 4), whereas the periodic
reviewed for fundamental understanding of decoupled combustion
component predominates in the deeper cavities (L/D > 4) and generates
processes and modeling in this section.
standing wave patterns. Shadowgraph imaging showed that the pressure
fluctuation was accompanied by the periodic shedding of vortices from
2.3.2. Flow oscillation mechanism
the front lip of the cavity and by acoustic radiation from the cavity with
Pressure oscillation in cavity flameholders may occur predominately
the principal source being close to the trailing edge of the cavity. The
in the spanwise, longitudinal or transverse directions under a wide range
experimental results showed that the principal characteristics of the
of flow conditions in supersonic flows [394]. Flow oscillation in cavities

39
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

sound waves in the cavity. A particular phase relationship can be


identified between the vortex shedding and the acoustic radiation. The
acoustic radiation reaches the front lip the cavity at the instant that a
vortex is shed, and a vortex is γ vλv behind the rear lip when this
particular phase of the acoustic radiation leaves the source at the rear
lip. There are mv complete wavelengths of the vortex motion and ma
complete wavelengths of the acoustic radiation involved. Fig. 49 shows
the simplified model of flow over the cavity. At time t = 0, an identified
phase of the acoustic radiation leaves the rear lip of the cavity and a
vortex is γ vλv behind the rear lip. At time t = t′ , an identified phase of the
acoustic radiation arrives at the front lip just as a vortex is shed. The
vortex pattern has moved downstream a distance kU∞t′ in this time in­
terval, so that

(9)

mv λv = L + γv λv + kU∞ t .
In the same time interval, the internal wave system has moved a
distance ct′ so that

(10)

L = ma λa + ct .
Eliminating t between the above two equations,

( )
kU∞ a 1
ma λa + (mv − λv )λv = kL M + , (11)
c c k
Fig. 49. Simplified model of longitudinal flow oscillation in a cavity [397].
and substituting for λv and λa,
periodic pressure fluctuation which occurred in the cavities may contain U∞ (ma + mv − γ v )
a number of periodic components, whose frequencies lie in a sequence f= ( ) . (12)
L
(m – γ) where m = 1, 2, 3 … and γ < 1. The frequency of any component M ac + 1k
was inversely proportional to cavity length and increased with free­
stream Mach number or velocity. Dimensional analysis suggested that Comparing this general equation to the empirical equation, the
the frequency of the longitudinal oscillation could be expressed in terms postulated interaction between the acoustic radiation and the vortex
of the Strouhal number (St) as a function of the freestream Mach number shedding is compatible with the experimental semi-empirical equation
(M∞) and the Reynolds number (Re), given that

L mv + ma = m ⎪ ⎪
St = f = f (M∞ , Re), (6) γv = γ

U∞ . (13)
k=K⎪ ⎪

where f is the longitudinal oscillation frequency, L is the cavity length, a=c
and U∞ is the freestream flow velocity. Based on the experimental It is noteworthy that the propagation speed of sound waves equals
observation and assumption above, a semi-empirical formula was the freestream speed of sound. Theoretically, the temperature in the
developed as cavity will be between the total temperature and static temperatures of
m − γ U∞ the freestream. However, the total and static temperatures differ about
fm = ⋅ , (7) 20% over the subsonic and transonic speeds range in these experiments.
M∞ + 1/K L
The use of the freestream temperature to calculate the speed of sound
where fm is the resonant frequency corresponding to the m-th mode, γ may involve an error of less than 10%.
and K are empirical constants. Physically, γ is the phase shift between In Rossiter’s model, the sound speed in the cavity is the sound speed
the acoustic waves and the shear layer instability, K is the ratio of the of the freestream, which is equivalent to assuming a cavity recovery
speed of the convection of the shear layer vortices to the freestream flow factor of zero. This assumption introduces a small error at low Mach
speed. The correlation constants used in Rossiter’s experiments were numbers, but the error is much greater at high Mach numbers. Heller
γ = 0.25 and K = 0.57. et al. [273,274,398] improved Rossiter’s formula for the high Mach
Rossiter [397] further provided a general physical model based on number range by assuming the cavity sound speed c to equal the stag­
the coupling between the acoustic radiation and the vortex shedding. It nation sound speed,
was assumed that the acoustic radiation initiated the vortex shedding [ ]1/2
and that the passage of the vortices over the rear lip of the cavity was γ − 1 2
c=a 1 + ∞ M , (14)
responsible for the acoustic radiation. The acoustic disturbances were 2
assumed to propagate upstream inside the cavity without disturbing the
the resonant frequency becomes
shear layer. The feedback loop is closed when the acoustic wave reaches
the upstream wall of the cavity, gives rise to a localized pressure force, m− γ U∞
fm = /√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ ⋅ , (15)
and excites the shear layer. The frequency of the acoustic radiation is the
M∞ 1 + [(γ ∞ − 1)/2]M ∞ + 1/K L 2
same as the vortex shedding frequency. The average speed of the
vortices over the cavity is k times of the freestream speed and the sound
where γ∞ is the freestream specific heat ratio. This formula can be
wave travels upstream in the cavity at a mean speed c, then
applied to shallow cavities with the length-to-depth ratio L/D ≥ 4. The
kU∞ c correction constants are γ = 0.25, K = 0.57 for cavities in airflows of
f= = , (8)
λv λa Mach numbers 2 and 3. The estimated error bound between the pre­
diction and experiments is ±10%. The accuracy of the prediction is
where λv is the spacing of the vortices and λa is the wavelength of the

40
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

equivalent to the motion of an infinite vortex sheet separating the


inviscid convecting flow with its velocity U+ from a quiescent inviscid
flow field of depth D. A periodic pressure pulse was adopted at the
leading edge of the cavity (the origin x = 0) to simulate the excitation of
the shear layer by the acoustic waves. A single periodic acoustic
monopole was used to model the dominant pressure fluctuations at the
trailing edge of the cavity (x = L) caused by the periodic mass inflow and
outflow. The pressure field from the trailing edge of the cavity was
assumed to have little effect on the vortex sheet except near the leading
edge. The leading and trailing edge processes were computed separately
and coupled with a phase shift of an integral multiple of 2π that selected
the allowable frequencies to close the feedback loop.
Fig. 50 shows the model of the driven shear layer spanning the cavity
and the image source distribution approximating the acoustic conditions
in the cavity. The stability of a vortex sheet subjected to a periodic
pressure impulse was studied at the leading edge of the cavity. The
asymptotic solution was obtained for the unsteady displacement on an
inviscid vortex sheet near a rigid boundary. The general form of the
steady state amplifying solution of the vortex sheet displacement func­
tion η(x,t) is represented for x > 0, by
η(x, t)∝exp(Ki x)exp[i(Kr x − ωt + ϕ)], (16)

where Ki and Kr are respectively the imaginary and real parts of complex
wave number in x direction, ω is the circular excitation frequency, ϕ is
the lag of vortex sheet displacement after forcing at x = 0. The flowfield
in the region of the trailing edge of the cavity was approximated using a
compressible simple source. The acoustic field was constructed by the
method of images and both the acoustic reflection from the vortex sheet
above the cavity and the pressure disturbances radiated by the oscil­
Fig. 50. Model of the driven shear layer spanning the cavity (top) and image lating sheet were neglected. The velocity potential approximating the
source distribution approximating the acoustic environment in shallow (mid­ cavity by an infinity image system is given by
dle) and deep (bottom) cavities [399].

1 ∑
∞ ( )
ω{ }1/2
Φ = Be− iωt
H0(1) [x − (2n + 1)L]2 + (y + 2mD)2 , (17)
limited as frequencies are also related to cavity configuration and flow m=0 n=− ∞
a−
parameters other than Mach number.
In the Rossiter and Heller models, the highly localized vortices are where H0 is the Hankel function of the first kind and zero order, B is a
(1)

the driving mechanisms for the shear layer instability and cavity oscil­ constant related to the mass flux, a- is the speed of sound inside the
lation. However, the vortex shedding may not be valid over the entire cavity (y < 0), m and n are summation indices in x, y direction, respec­
Mach number range as far as cavity oscillation are concerned [348]. tively. The pressure field is obtained from
Experimental visualization showed that no vortex shedding was found at
certain cavity configurations even though intense oscillations were P = iωρ− Φ. (18)
produced between the fluid in the cavity and the free shear layer
spanning the cavity mouth when the external flow ranged from high
ρ- is the density inside the cavity volume. The pressure at the leading
edge of the cavity is obtained by setting x = y = 0, approximating Hankel
subsonic to supersonic speeds [274]. Rossiter’s and Heller’s
functions of large argument with asymptotic representations, and
semi-empirical formulas can predict experimental results within its
neglecting all terms but the source at the trailing edge,
applicable range and it does not necessarily indicate that these models
are physically sound. Although the Rossiter model is developed based on
the coupling between the acoustic radiation and the vortex shedding, it
is not described how acoustic disturbances are generated at the cavity
trailing edge and how the feedback acoustic waves excite the shear layer
at the cavity leading edge.
Bilanin and Covert [399] improved the Rossiter model by relating
the driving mechanism of cavity oscillations to the instabilities of the
free shear layer. In the Bilanin model, the flow instability waves of the
shear layer are periodically excited at the cavity leading edge and grow
as they propagate downstream. The fluctuating motion of the shear layer
at the cavity trailing edge results in the periodic mass flow exchange
between the cavity and freestream, i.e., an entrainment of the external
flow into the cavity within half a period and a discharge of cavity flow
into the external flow within another half a period. The periodic action
of the mass inflow and outflow is the source of the acoustic radiation,
which is different to that in the Rossiter model. The analytical model was
developed based on the assumption that the shear layer was a thin
vortex sheet and its motion was more strongly dependent upon the
bottom boundary than the ends. The shear layer dynamics was
Fig. 51. Acoustic wave field inside and outside a rectangular cavity [348].

41
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

( )
ωL whereas the acoustic waves radiated into the cavity are reflected by the
P = 2iωρ− Be− iωt
H0(1) . (19)
a− bottom wall and the upstream end walls of the cavity. Nevertheless, it
was shown that the secondary reflections from the bottom wall of the
The mass flux from the compressible source is
cavity had little effect on the excited shear layer instability for cavity
( )
ω with aspect ratio L/D < 2 and the amplitude of the instability wave
ṁ = 4i Be− iωt . (20) excited by the directly reflected waves from bottom wall at cavities with
a−
aspect ratio L/D > 2 was small compared with the original excited
The mass flux into the cavity near the trailing edge is approximately instability waves at the leading edge of the cavity. The presence of the
ṁ ∝ − η(L, t), (21) bottom cavity wall was ignored on modeling the instability character­
istics of the shear layer.
and is the monopole strength. By using these results to eliminate the In the Tam model shown in Fig. 51, the acoustic wave field from
source strength, the formula for the pressure at the leading edge of the unsteady interaction between the shear layer and the trailing edge of the
cavity is obtained cavity is simulated by a first-order imaginary periodic line source Q
placed at the trailing edge. The directly radiated wave EE and the re­

( )
ωL
P∝ei[Kr L− ωt+ϕ+π] H0(1) . (22) flected waves A A and B B all tend to excite the instability waves of the
′ ′′ ′ ′′

a−
shear layer. The wave field B B′′ can be simulated by using a periodic line

The pressure is to force the vortex sheet at x = 0 and have the same image source B located at x = L, ​ y = 2D, while the reflected wave field
the phase as e-iωt. The eigenvalue equation for excitation frequencies is A A′′ can be simulated by means of an image line source A at x = −

obtained by requiring L, ​ y = 0. The secondary reflections are ignored in the model. The
3 ωL excitation of unstable waves of a turbulent shear layer by sound is
Kr L + ϕ + π + = 2nπ, (23) modeled within the framework of hydrodynamic stability theory from
4 a−
Chan [400–402] and Moore [403]. The vortex sheet displacement at the
where n = 1, 2, 3, …. It is equivalent to the Rossiter’s semi-empirical trailing edge of the cavity is represented by
model in the form
ξ(L, t) = iQκ exp(− iωt), (25)
L n− γ
St = f = (24)
U+ M+ a+ /a− + 1/K where Q is the line source strength and κ is a function of Strouhal
number ωL/U∞ , Mach number M∞ , length to depth ratio of the cavity
with γ = − (3 /8 +ϕ /2π) and K = ω/Kr U+ . This improved model is also L/D, and the momentum thickness θ,
consistent to the modified model from Heller et al. [273]. ( / )
In the Bilanin model, the process of the acoustic disturbance inter­ κ = ψ ωL U ∞ , M, L / θ, L / D , (26)
acting and exciting the flow instability of a thin shear layer was not
considered. This model showed considerable discrepancy between the
where ψ is a known complex-variable expression from the hydrody­
theory and experiment results from Rossiter at Mach number 0.6–1.2,
namic stability theory analysis. The left-hand side of the above equation
nevertheless, the predicted excitation frequency agreed well with mea­
is real; in order to close the feedback loop, the phase of ψ must be equal
surement in supersonic flows with Mach number up to 3.4 from Heller
to an integral multiple of 2π, i.e.,
et al. It had been shown that the infinitesimally thin vortex sheet became
stable at a sufficiently high Mach number [348]. At high Mach number phase ​ of ​ ψ = 2πn; ​ n = 1, ​ 2, ​ 3, ​ … ​ . (27)
when the vortex sheet is stable, the driving mechanism of shear layer
instability for cavity oscillation vanishes in the Bilanin model, which where n is the order of the oscillatory mode. The Tam model has overall
contradicts experimental observations. This change in the instability good agreement between the predicted discrete tone frequencies of the
characteristics of a thin shear layer was not evaluated in the Bilanin proposed model and the pressure oscillation data from Rossiter at Mach
model although it could be applied to high Mach number flows. number 0.4–1.2 and the measured data from Tam at Mach number
Tam and Block [348] further extended the Bilanin model by ac­ 0.2–0.4.
commodating the features of the shear layer instability and acoustic
wave generation process by the impingement of the shear layer near the
trailing edge. The detailed acoustic wave generation mechanism was
proposed that differed to the unsteady mass addition and removal pro­
cesses in the Bilanin model. Fig. 51 shows the acoustic wave field
induced by the unsteady shear layer over a rectangular cavity. This
model accounts for the finite shear layer thickness effects and the shear
layer was simulated using a vortex sheet.
In the Tam model, the acoustic wave is generated by the shear layer
oscillation up and down near the trailing edge of the cavity. The fluid of
the shear layer shields the trailing edge of the cavity from the external
flow during the upward motion of the shear layer, which smooths the
external fluid flowing over the trailing edge and minimizes the gener­
ation of pressure waves of any significance in this period of a cycle. The
inflow of external fluid into the cavity is formed at the trailing edge
when the shear layer is deflecting downward and a momentarily high-
pressure region is generated near the trailing edge of the cavity.
Compression waves develop at the trailing edge of the cavity and
propagate in all directions. The shape of the acoustic wave front can be
modified by the convection effect of the mean flow during the propa­
gation away from the trailing edge of the cavity. The acoustic waves
Fig. 52. Schematics of fuel injector and spark plug configurations within/near
radiated into the external flow propagates to infinity without reflection, a cavity flameholder [406].

42
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

2.3.3. Ignition process in flameholders 1530 K, 2.6 MPa, and 2.92, respectively. The schematics of the fuel in­
Ignition in a scramjet combustor is challenging considering the jectors, spark plug locations and the cavity dimensions are presented in
complicated fuel-air mixing process and the low pressure during the Fig. 52. The parallel injection upstream of the cavity provided a wider
engine start-up stage. The critical issue regarding a reliable fuel ignition region of overall equivalence ratio and spark plug location for successful
scheme in scramjet combustors is the creation of a subsonic combustion ignition than the cascading injection scheme upstream of the cavity. The
region. As a simple cavity is used as an integrated fuel injection and spark ignition location near the cavity rear wall demonstrated the best
flameholding device, ignition devices have been widely used inside the ignition performance, hypothetically due to the favorable ignition
cavity volume due to the relatively well-mixed combustion reactants, environment provided by the chemical reaction and flow conditions.
low-speed flow recirculation, and heat recovery. The ignition process is However, the flow conditions favorable for the ignition process are not
highly sensitive to fuel injector location, fueling rate, cavity geometry, reported, and the interaction between the chemical reaction and flow
and location of ignition kernel. Ignition sources include spark plugs conditions is not clearly identified.
[404–407], laser induced plasmas [408–410], and pulse detonators The transient ignition processes of liquid hydrocarbon fuels are more
[407]. challenging during the engine startup stage due to the fuel addition
Sun et al. [404] studied the spark ignition process in a processes for ignition, e.g., atomization, evaporation, and the increased
hydrogen-fueled scramjet combustor at Mach 4 free flight condition. A ignition delay for large molecules. For instance, the flame development
direct-connect combustion-vitiated test facility was used to generate the processes of the vaporized kerosene may fail at different ignition stages
Mach 1.92 freestream at a stagnation temperature of 846 K, and a without proper initial fuel concentration distribution, which is related to
stagnation pressure of 0.7 MPa. A spark plug was flush mounted at the the injection pressure [411]. Usually the combined injection of gaseous
center of the cavity bottom wall. Hydrogen fuel was transversely and liquid fuels is used to broaden the fueling conditions for successful
injected into the crossflow upstream of the cavity. It was observed that spark ignition and global flame stabilization of the integrated
the flame kernel was initiated near the hot spark plasma region, spread ignition-flameholding scheme. Hydrogen fuel has been used to ignite
along the cavity shear layer, and ignited the fuel distributed down­ and pilot kerosene flames.
stream. With the cavity pressure buildup during the ignition transition Xi et al. [411] studied the piloted ignition of an ethylene flame for
process, the pre-combustion shock trains and the flame region moved kerosene fuel in a direct-connect combustor. The airflow at the
upstream until the flame was stabilized surrounding the jet. The ignition combustor inlet had a core Mach number of 2.52, a stagnation pressure
performance was observed to be significantly increased by direct of 1.6 MPa, and a stagnation temperature of 1486 K. The fueling range of
hydrogen fuel injection into the cavity ramp region. No detailed flame upstream kerosene injection for successful spark ignition inside the
structures were reported during the transient ignition process. cavity was extended by the piloting an ethylene flame stabilized be­
Cai et al. [406] investigated the effects of fueling schemes and spark tween the kerosene jet and the cavity leading edge. The ignition limi­
plug locations on the ignition process in an ethylene-fueled scramjet tations were broadened by the intensified mixing and evaporation
combustor. The airflow at the direct-connect combustor inlet had a processes as a result of the high-temperature heating of the ethylene
stagnation temperature, a stagnation pressure, and a Mach number of flame and the pre-combustion shock train driven by the increased

Fig. 53. High-speed broadband chemiluminescence images of transient ignition and flame stabilization processes with different fueling configurations [405].

43
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 54. Shadowgraph and chemiluminescence images of cavity ignition by spark discharge and pulse detonator [407].

aerodynamic blockage. Detailed mechanisms of intensified mixing and pulse detonator was more effective for ignition at lower flow tempera­
evaporation processes by the pilot ethylene flame were not further tures and pressures by providing high pressure and temperature exhaust
quantified. and a significant disruption to the flow. It was shown however that the
Tian et al. [129,405] investigated the effects of the pilot hydrogen pulse detonator may decrease cavity burning and even cause cavity
flame location on ignition and flame stabilization in a kerosene-fueled extinction under certain cavity fueling conditions during the ignition
scramjet combustor at Mach 4.5 flight condition. The combustor had process. The strong initial flame kernel can increase the cavity pressure
inflow conditions of Mach 2.0, a total pressure of 1.0 MPa, and a total rapidly, which leads to the blockage of the cavity fuel and starvation of
temperature of 1100 K. Fig. 53 presents the transient ignition and flame the cavity until the pulse detonator is completely exhausted. The
stabilization processes with various fueling schemes. The kerosene decrease in burning during the pulse detonator ignition process could be
injected upstream of the cavity can be successfully ignited by the pilot mitigated with sufficiently high cavity fueling.
flame generated by direct hydrogen injection into the cavity, but the The strut-type flame holder is also widely applied in scramjet engines
kerosene flame failed to self-sustain when the pilot flame was removed. though strut is commonly used as a fuel injector as reviewed in the
The kerosene injected into the cavity demanded increased hydrogen section of mixing enhancement methods. This type of flameholder
fueling rate upstream of the cavity for successful ignition. The kerosene typically operates in the lower hypersonic flight regime because the total
flame could be self-sustained after successful ignition. It was concluded pressure loss introduced by struts can be tremendous at increased Mach
that direct cavity injection of liquid kerosene generated high levels of numbers. The ignition process of a kerosene-fueled supersonic
heat release and sustained the combustion within a blowout limit after combustor with strut injection is shown in Fig. 55 [414–416]. The initial
the removal of the pilot hydrogen flame. flame appears around the ignition source in the recirculation zone at the
In recent years, laser induced plasma (LIP) has been considered to be trailing edge of the strut and grows to form a steady flame in the core
a promising ignition technique in practical propulsion systems. The LIP flow. Different flame establishment processes were found for varying
ignition scheme has many advantages compared to the traditional spark equivalence ratios. At the low equivalence ratio, the initial flame ap­
plug ignition, such as the flexibility of ignition location, controllability pears far downstream of the strut, and then propagates upstream. A
of ignition timing and energy deposition, and immunity from flowfield liftoff region develops due to the insufficient mixing of fuel and air. At
interference. The LIP can be placed in a fuel-air mixing region favorable high equivalence ratio, the initial flame establishes in the recirculation
for ignition that is far away from the combustor wall. Accurate controls region formed at the strut end, and spreads into the mainstream in
on the LIP generation time and energy level facilitate studies on tran­ streamwise direction. An amount of oxygen can be injected into the strut
sient ignition phenomena and ignition probability in high-speed flow­ wake region through orifices a the back of the strut to promote ignition
fields. The fundamental physical process of laser induced spark ignition and form a stable central local flame [416–418]. The thickness of strut
was reviewed by Phuoc [412], and its application in aerospace propul­ can be reduced by oxygen injection without affecting the ignition and
sion was reviewed by O’Briant et al. [413]. Recently, much progress has flameholding ability. Drag and total pressure can be minimized by
been made in the application of LIP to fuel ignition inside cavity reducing the thickness of the strut. Due to the harsh heating environ­
flameholders in scramjet engines. A summary of laser ignition inside a ment encountered by strut flameholders, typically liquid hydrocarbon or
cavity will be presented in the section on plasma assisted combustion. hydrogen fuels are injected at the leading edge or the front section of the
Alternative energy deposition techniques, such as pulsed detonators strut for thermal protection.
and plasma torches, have been used for fuel ignition and flame stabili­
zation purposes in supersonic combustors. The ignition of an ethylene 2.3.4. Combustion stabilization mechanism
fueled cavity in supersonic flow by a spark discharge and a pulse deto­ Combustion stabilization by cavities needs to be achieved at a variety
nator as shown in Fig. 54 [407]. The spark discharge ignition was pas­ of flow conditions and fueling schemes, taking into consideration the
sive with the ignition kernel and ensuing the flame propagation dependence on freestream Mach number, flow stagnation temperature,
following the cavity flowfield. Compared to the spark discharge, the fuel injection method, cavity geometry, etc. Cavity flameholding can be

44
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 55. Ignition process of a kerosene-fueled supersonic combustor installed with a simple strut [407].

categorized based on the flame luminosity distribution and flame (1390–1490 K), the jet wake stabilized flame structures were consistent
structures under passive and active fueling schemes in steady ram- and with a lifted jet flame, in which three flame regions were observed, i.e.,
scram-mode operations. Combustion stabilization mechanisms are the liftoff region, a thickened premixed flame region at the lifted flame
determined by combined effects from the strength of the recirculation base, and the turbulent non-premixed flame zone with highly thickened
zone and the local stoichiometry determined by fuel-air mixing pattern and shredded flamelet structures. In the lower flow stagnation temper­
[419–422]. ature range (1200–1300 K), in which the fuel cannot be auto ignited, the
Passive cavity flameholding configurations inject fuels upstream of reaction zone has an averaged constant spreading angle and a relatively
recessed cavities. Fuel-air mixtures are entrained into the cavity and the continuous reaction layer emanating from the top of the cavity leading
stabilized flames spread into the main flow from the cavity. Micka et al. edge. These features are consistent with the flame spreading controlled
[423–427] characterized cavity-stabilized combustion in a reaction of a premixed flame stabilized by the cavity recirculation zone.
hydrogen/ethylene-fueled dual-mode scramjet combustor. The isolator A hybrid flame stabilization mechanism termed autoignition-assisted
airflow has an entrance Mach number of 2.2 and a total temperature of flame was proposed based on the flame reaction zone structures for
1040–1490 K corresponding to a flight Mach number of 4.3–5.4. The both combustion modes.
combustor consists of a wall-normal sonic fuel jet injected into a su­ Micka et al. [424] observed that only a cavity-stabilized combustion
personic crossflow upstream of a cavity flameholder. Fig. 56 shows the mode was achievable in scram-mode operation though the flow stag­
two distinct combustion stabilization locations identified in ram-mode nation temperature was high enough to auto ignite the fuel. The high
operation. The flame base was anchored at the leading edge of the velocity and low static temperatures associated with a supersonic flow at
cavity by heat release in the cavity shear layer at low flow stagnation the isolator exit made the jet-wake location unsuitable for flame stabi­
temperatures and stabilized a short distance downstream of the fuel jet lization in the combustor. Nevertheless, both jet-wake-stabilized com­
in the jet wake region. In the higher stagnation temperature range bustion and cavity-stabilized combustion have been observed in a

45
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 56. Combustion stabilization modes: (a) jet wake stabilized flame at higher flow stagnation temperature, and (b) cavity stabilized flame at lower flow stagnation
temperature [424,427].

supersonic combustor with similar flow conditions but with different recirculation zone stabilized flame regime. In the shear-layer stabilized
jet-cavity configurations [428,429]. In scram-mode operation, the flame regime, the flame base anchored downstream of the cavity front
combustion stabilization mode can be modified by the overall equiva­ step and stayed in the shear layer where the flame velocity was equal to
lence ratio, the distance from the fuel jet to the cavity leading edge, the the air velocity. A lifted non-premixed flame was located along the
cavity length-to-depth ratio and the ramp close-out angle. stoichiometric contour. In the recirculation zone stabilized flame
Wang et al. [428,429] studied the combustion characteristics of a regime, the flame base attached to the front step of the cavity and the
hydrogen jet upstream of a cavity flameholder in a supersonic flame zone spread into the cavity volume, which lifted off the shear layer
combustor. The freestream at the combustor inlet had a Mach number of and increased the fuel-air mixing and flow entrainment into the cavity.
2.52, a stagnation temperature of 1486 K, and a stagnation pressure of O’Byrne et al. [431–433] investigated the combustion stabilization
1.6 MPa to simulate free-jet flight conditions at Mach 6. Three com­ in a cavity-based supersonic combustor in high-enthalpy flow conditions
bustion modes in scram-mode operation were identified: cavity assisted (6.45–3.82 MJ/kg). A free-piston shock tunnel was used to generate
jet-wake stabilized combustion, cavity shear-layer stabilized combus­ realistic combustor entrance conditions for flight at Mach 11.4 at an
tion, and the combined cavity shear-layer/recirculation stabilized altitude of 29 km. The flow at the combustor entrance had a nominal
combustion, as shown in Fig. 57. A short cavity installed closely Mach number of 4. The hydrogen fuel was injected through four sonic
downstream of the jet can assist the jet-wake stabilized combustion, nozzles at a 15◦ inclination toward the open cavity. The OH fluorescence
which was unstable and experienced intermittent blowout. Combustion image taken at ϕ = 0.44 is shown in Fig. 58. A low equivalence ratio did
stabilization in large cavities depends on the distance between the jet not ignite at the jet location, but at higher equivalence ratios the flow
and the cavity. In combined cavity shear-layer/recirculation stabilized started to ignite immediately upstream of the injector. A measurable
combustion, strong reactions occurred in both the cavity shear layer and amount of OH radicals was observed within the cavity over the entire
recirculation region, and a large amount of heat was released in and range of equivalence ratios, which verified the cavity flameholding
around the cavity. In cavity shear-layer stabilized combustion, the flame ability of this fueling configuration though most of the combustion was
base anchored at the cavity leading edge and the reaction zones still stabilized in the shear layer above the cavity. Additional pressure
appeared only in the cavity shear layer with heat release mainly and fluorescence measurements showed that the heat release was mostly
occurring in the external region of the cavity. Increasing the overall initiated by the shockwave from the cavity trailing edge, and the igni­
equivalence ratio or increasing the cavity close-out/aft angle resulted in tion above the cavity did not have a strong influence on the downstream
the transition from cavity shear-layer stabilized combustion to com­ combustion. The cavity’s impact on the overall heat release was sec­
bined cavity shear-layer/recirculation stabilized combustion. The com­ ondary to the effect of the oblique shock generated by the rear face of the
bustion spreading from cavity shear layer to freestream was proposed to cavity at high-enthalpy flow conditions.
be dominated by both diffusion and convection processes associated Liu et al. [434] investigated the dominant flameholding mechanism
with the extended recirculation flows resulting from the heat release and and the effect of turbulence on combustion for identical flow and cavity
the interaction between the jet and the cavity shear layer. configurations using LES. Fig. 59 presents the typical distribution of fuel
Le et al. [430] investigated the combustion stabilization of liquid and combustion products in the contours of density gradients. Mass flow
kerosene fuel by a downstream cavity flameholder. The Mach number, exchange between the cavity and the jet wake region was found to be
stagnation temperature and stagnation pressure of the airflow were 2.6, essential for flameholding. The intermediate high-temperature gases
1490 K, and 1.7 MPa, respectively, which are similar incoming flow generated inside the cavity provided heat continuously to ignite the
conditions to Wang et al. [428,429]. Two flame stabilization modes fuel-air mixture in the jet wake, which stabilized the flame in the cavity
were defined, i.e., the shear-layer stabilized flame regime and the shear layer. Two combustion modes coexisted in the global reacting

46
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 57. Combustion modes in scram-mode operation. Left: OH flame chemiluminescence images; right: schematics of stabilization mechanisms [428,429].

419–422]. In the case of direct fueling into recirculation region, the


reactants entering the stabilized combustion regions are non-premixed,
hence the flame stabilization limit and flame structures deviate from the
cases in cavity flameholding with upstream fuel injection. The flame
stabilization in different cavity regions depends on the strength of the
recirculation zone, the local stoichiometry, and the flow pattern created
by the fuel injection location/direction. In addition, active fuel injection
into the cavity can generate a pilot flame to stabilize the combustion
downstream of the cavity.
Rasmussen et al. [419–421] systematically surveyed the flame­
Fig. 58. An OH-PLIF image of flame at freestream total enthalpy h0 = 6.45 MJ holding characteristics of a rectangular cavity with direct fuel injection.
and overall equivalence ratio ϕ = 0.44 [433]. The Mach 2 freestream had a stagnation temperature of 590 K and a
stagnation pressure of 5.44 atm. Fuel injector location was observed to
regions: the steady diffusion flame and the unsteady flame branch. The have a large effect on the cavity flameholding mechanism in the directly
steady turbulent diffusion flame dominated the combustion downstream fueled cavity flameholder, as shown in Fig. 60. When the fuel was
of the cavity with major heat release occurring near the stoichiometric injected from the aft wall, the fuel immediately entered a combustion
mixture fraction. product region between the shear layer and bottom wall, recirculated
Direct fuel injection into the cavity provides a wide range of flame along the cavity floor, and combusted along the underside of the shear
stabilization over passive fuel injection upstream of the cavity. The layer. When the fuel was injected from the bottom wall, a jet-driven
directly fueled cavity can maintain combustion stability during the recirculation zone developed at the upstream wall of the cavity, which
transient ignition process, in which the scramjet operation mode tran­ provided a radical reservoir to reduce the combustion ignition delay
sitions between the scram- and ram-modes as a result of flight Mach under the shear layer. For both fuel injection methods, the reaction
number and combustion heat addition into the core flow [106, anchored in the shear layer at moderate fueling conditions. Near lean or

47
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 59. Pressure on the bottom wall and the subgrid turbulent kinetic energy in the streamwise slice associated with 3D streamlines [434].

Fig. 60. Combustion distribution (left: CH2O PLIF images) and stabilization mechanisms (right: schematics) of cavity-fueled flameholding scheme at fuel-lean and
fuel-moderate conditions with wall injection and floor injection [421].

rich blowout limits, the reaction zone stabilized differently for wall and shear-layer/recirculation mode achieved by the direct cavity injection,
floor injections. but enlarged the flame spreading angle from the cavity to the external
Combined upstream and cavity injections can further increase the flow.
steadiness of cavity-stabilized combustion. Micka et al. [424] reported Flameholding characteristics of strut-type flameholders has been
that the rear wall and floor fuel injections inside a cavity could extend widely investigated in terms of the ignition, piloting, propagation, and
the reaction zone of jet-wake stabilized combustion into the upstream stabilization. Combustion stabilization characteristics of strut flame­
part of the cavity, i.e., a hybrid mechanism with primarily jet-wake holders can be impacted by the properties of injected fuels (hydrogen
stabilization and secondarily cavity-assisted stabilization. This hybrid fuel [250,308,435–440], hydrocarbon fuels [372,378,414,415,418]),
stabilization was more steady than pure jet-wake stabilization but less piloting configuration [390,416,417,441,442], strut geometry [308,
steady than pure cavity stabilization. Wang et al. [428] observed that 372], strut jet diameter [372], downstream installation of staged strut
combined fuel injection did not modify the cavity [389,391–393,443,444], step [445], cavity [378,379,381,383,385,386,

48
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 61. Contours of temperature, heat release rate, diffusion flame index, the premixed flame index in strut-stabilized hydrogen flames [459].

446–448], ramp [449], and secondary jets [415,450], and the secondary understanding of the dominant parameters affecting overall supersonic
flow structures created by shaped combustor walls [451]. Streamwise combustor performances, a simple strut is typically installed in a planar
vortex generators (e.g., 3D jets and alternating wedges) can be inte­ two-dimensional combustor with area relief to study the large-scale
grated to strut fuel injectors to promote the near-field mixing as combustion characteristics as well as local flame stabilization at the
reviewed in the section of mixing enhancement methods. For better strut recirculation zone and wake region. Substantial studies on

Fig. 62. Flame stabilization modes as a function of inlet stagnation temperature T0 and fuel-equivalence ratio ϕ (left) and shearing interferogram of the flame [479].

49
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

strut-based combustors use the well-documented experimental database The coupled dynamics between high-speed flow and combustion are
from DLR [452,453] as validation benchmark over the last 20 years established when acoustic waves propagate upstream in subsonic flow
[451,454–469]. Another experimental database of strut-stabilized su­ regions, which closes the feedback loop of fluid-combustion interaction.
personic combustion from JAXA has different strut and combustor ge­ Lin et al. [473] developed an analytical model of thermal acoustic in­
ometries [308,309,438] to the DLR combustor. It has also been widely stabilities of the acoustic and convective feedback loops among the
used to validate the numerical models of supersonic combustion and to pre-combustion shock, fuel injection, and distribution of flames in
study strut-based flameholding mechanisms [470–472]. ramjet combustors. Wang et al. [96,474–478] studied and summarized
Fig. 61 shows the distributions of temperature, heat release rate, the physical mechanisms that lead to combustion oscillations at low and
flame index in strut-stabilized hydrogen flames [459]. The stabilized high frequency regimes in scramjet combustors. In general, the
flames can be divided into three regions. In the first region where the high-frequency oscillations are related to self-sustained cavity oscilla­
fuel is discharged into the wake of the flameholder at the strut end, tion, cavity shear layer instabilities, and flow/flame unsteadiness
diffusion combustion mostly exists in the shear layers shed off the strut resulting from shock-boundary layer and flow-flame interactions. These
edges, and the heat release rate is small. The second region starts from oscillations are relatively mild and less prone to result in combustion
around a characteristic length of strut downstream of the wake, in which extinction. Combustion instabilities associated with combustion mode
diffusive and premixed combustion produces significant heat release. As transition during cavity flameholding are typically more violent and
the high flow speed and the short residence time limit the mixing be­ severe due to the tendency toward cavity flame blowout. Combustion
tween hot combustion products and freestream, the high-temperature oscillations coupled with combustion mode transition are summarized
zone is narrowed. In the third region starting from around four times in this section.
of the characteristic length of strut, turbulent mixing and lean Micka et al. [424] and Fotia et al. [479] studied ram-scram mode
post-combustion succeed with significant heat transfer. In this region, transition in a dual-mode scramjet combustor and observed combustion
diffusive combustion dominates at the edges of vortex structures and oscillation existing between the cavity stabilized ram-mode at low
premixed combustion occurs in the cores of the vortex structures. The stagnation temperatures and the jet-wake stabilized ram-mode at high
high-temperature zone is broadened and the heat release is weak. stagnation temperatures. Fig. 62 shows the combustor operation regime
and the flame stabilization mode as a function of inlet stagnation tem­
2.3.5. Combustion mode transition and oscillation perature and overall equivalence ratio. In the case of unsteady flame
Combustion instabilities occurring in scramjet combustors are stabilization with low-frequency periodic oscillations (~5 Hz) in
related to the intrinsic flow instabilities. The flow instabilities may result ram-mode operation, flames were observed in the forward jet-wake
from self-sustained cavity flow oscillations, flow unsteadiness from jets stabilized location, the rearward cavity-stabilized location, and the in­
in supersonic crossflow, shockwaves, and separated boundary layers. termediate lifted jet location. The entire shock train underwent

Fig. 63. Schematic of physical mechanism driving the scram-to-ram mode transition [383,480].

50
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

oscillation at the same frequency but with a different phase angle. A


mechanism was identified, i.e., the flame speed excited the
self-sustaining cavity shear layer instability and drove the oscillatory
behavior in the unstable regime.
Aguilera and Yu [383,480] investigated the scramjet to ramjet
transition in a dual-mode combustor with fin-guided injection upstream
of the cavity flameholder. The physical mechanism of oscillating flame
behavior observed during the scram-to-ram transition is summarized in
Fig. 63. Initially the combustor operated in scramjet mode with flame
anchoring downstream of the cavity. The flame propagated upstream
accompanied by a shock train due to transient thermal choking. The
flame failed to remain attached inside the cavity, detached from the
ignitor, and was pushed back into the diverging duct. The flame reces­
sion caused the shock train upstream of the fin to be swallowed and
reinitialized scramjet operation. The increase in flow heat addition in
the downstream portion of the combustor increased the frequency of
upstream flame propagation and flame presence in the cavity. The
transition of combustion to a scram mode where the flame was blown
downstream occurred less often when the cavity was heated up which Fig. 65. Flow oscillation frequencies obtained from experimental measure­
also led to less flame quenching. The scram-to-ram mode transition in a ments and analytical predictions based on various feedback loops [473].
combustor with a fin-cavity configuration occurred with a series of mode
hopping events that evolved over a period of time related to the fueling acoustic feedback, shock-flame acoustic-convection feedback, and the
condition. injector-flame feedback. The shock-flame acoustic feedback is charac­
Yentsch et al. [135–137,481] numerically investigated the unsteady terized by the admittance function for a normal shock in a diverging
scramjet mode transition phenomena in the HIFiRE-2 flowpath. It was channel, and was developed to predict the shock oscillation driven by
reported that the separation associated with unsteady shock/turbulent downstream combustion [473,491,492]. The characteristic times of
boundary layer interactions was critical during mode transition, and low purely acoustic and acoustic-convective feedback loops between the
frequency oscillations were observed to be present in the dual- and shock and flame zone were estimated as below,
scram-mode operation.
Air throttling is considered as an effective method for scramjet /
fsf 1 = 1 τsf 1 ; τsf 1 ≃ (
2Lsf
), (28)
ignition and flame stabilization [129,482–490]. Air throttling can a 1 − M2
generate a pre-combustion shock train to initiate combustion in a
scramjet combustor, and the shock train can be retained leading to
/ L
sustained combustion if sufficient heat release is produced in the fsf 2 = 1 τsf 2 ; τsf 2 ≃ ( sf ), (29)
combustor. Yang et al. [482] investigated the air-throttling process in a aM 1 − M
supersonic combustor with a total temperature and a total pressure of
953 K and 0.82 MPa, respectively. It was observed that the flow was where a and M represent the speed of sound and Mach number longi­
oscillating when the shock front was positioned between the cavity front tudinally averaged between the shock and flame, respectively, and xs
step and rearward step during ram mode operation after flow choking and xf are the locations of the shock and flame, respectively. The
was induced by air throttling. The self-sustained oscillation was related
to the periodic deflection of the shear layer into the main flow driven by
the high back pressure in the cavity separation zone. Tian et al. [129,
488,490] explored the oscillatory behaviors and the pseudo shock dy­
namics during the mode transition process in a kerosene fueled
combustor with the same experimental configurations and flow
conditions.
Mechanisms of combustion oscillation in ramjet and dual-mode
scramjet operating in ram mode have been developed. Acoustic feed­
back loops are considered only in the subsonic region bounded by the
pre-combustion shock in the isolator and the thermal throat in the
downstream region of the flame zone [473]. Fig. 64 shows three acoustic
feedback mechanisms which have been identified, i.e., the shock-flame

Fig. 64. Acoustic-convective feedback loops and associated characteristic ve­ Fig. 66. Illustration of a perfectly stirred reactor simulation of a mixing shear
locities in a scramjet combustor [473]. layer and cavity dimensions [500].

51
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

oscillating heat release caused by the flow-rate oscillation in the fuel dominant factor that determines the overall blowout limit [365]. Gruber
injection/mixing region was correlated to the acoustic oscillation, [420] provided an overall assessment on the flame stability with passive
fuel entrainment and direct cavity fuel injection. It was observed that the
Lif
τa ≃ , (30) direct injection from the cavity ramp was more desirable than the pas­
a− u
sive entrainment in many aspects, which includes a wide range of fuel
Lif flow rate for stable cavity combustion, the uniform distribution of
τc ≃ . (31) fuel-air mixture within the cavity, and the relative insensitivity to
u
external flowfield changes. Donohue [493] found that the blowout
The characteristic frequency for the acoustic-convective feedback loop temperature limit of cavity flameholding for a dual-mode scramjet was
between the fuel injector and the flame zone is most sensitive to the fuel injection location (external and internal cavity
⎡ ⎤− 1 fuel injection). Lin et al. [495] observed a significant variation in shape
/ L and spatial distribution of the cavity flame at various fuel flow rates and
(32)
if
fif = 1 τif = 1 / (τa + τc ) ≈ ⎣ ( )⎦ . cavity fueling methods.
aM 1 − M
In general, cavity flammability limits are correlated with the flow
entrainment rate from the freestream into the cavity. Characteristic
The experimentally measured and analytically predicted oscillation
airflow rates of recessed cavities in rectangular combustors were
frequencies are compared in Fig. 65. Nearly all the measured oscillation
assumed to be scalable with the freestream mass flux and cavity opening
frequencies fall in the range stipulated by the shock-flame acoustic-
flux area [106]. In all scaling practices, the lean and rich blowout limits
convective and the injector-flame feedback loops. Both the shock-flame
in terms of the direct fueling mass flow rates and the characteristic air
and injector-flame instability mechanisms were observed in most ex­
flow rate at fixed freestream conditions were close to a linear correlation
periments, and the corresponding frequencies match the predictions.
[419,420], which proves the scaling relation between the cavity
The predicted frequencies have moderate uncertainties due to the sim­
entrainment mass flow rate and the characteristic air flow rate. It is
plifications adopted in the analytical model.
generally accepted that cavities do not scale in size mainly due to the
fact that chemical ignition delay requires a certain time to mix. The
2.3.6. Flame stability limit
impacting factors showed to significantly affect the cavity flow
Flame stability limit of non-premixed combustion is the critical
entrainment include cavity geometry (length-to-depth ratio, ramp
parameter for cavity flameholding design in dual-mode scramjet engines
[106]. Flame stability limit in a cavity depends on the freestream and angle), fueling schemes, and freestream conditions (Mach number,
stagnation pressure and temperature).
fueling conditions, which include freestream stagnation temperature,
The operating limits of recessed cavity flameholders with various
Mach number, fuel type, fuel injection scheme, and cavity geometry.
cavity geometries were reported in several studies. Lin et al. [496,502]
Most studies on flammability limit [380,420–422,430,493–499] were
showed that the cavity with larger length-to-depth ratio had poorer
focused on factors that impact blowout limits (freestream Mach number,
performance in terms of lean operability but the associated flowpath
stagnation temperature, equivalence ratio, Damkohler number, com­
performed better in terms of overall combustor operability and thrust
bustion contamination effect) on a case to case basis, and very few
generation. Rasmussen et al. [419,420] observed that cavity lean and
theoretical models [106,500,501] were developed due to the intrinsic
rich blowout limits were affected by the configuration of cavity closeout
complexity of the cavity flameholding problem.
angle. Floor injection from a rectangular cavity (no ramp) resulted in
Davis and Bowersox [500,501] developed a perfectly stirred reactor
lower lean blowout fuel flow rates than the case of a cavity with a
(PSR) model for cavity flameholders to analyze effects of inflow tem­
slanted ramp. The ramp angle controlled the location where the shear
perature, heat loss, trace species diffusion, overall equivalence ratio, and
layer impinged on the cavity rear wall and hence affected the air
fuel type on cavity flammability. The schematics of the PSR model are
entrainment rate.
shown in Fig. 66. The ratio of products and reactants entering the shear
Fuel injection direction and location were proved to be important
layer is generally unknown. The reactor volume increases in time to
factors affecting the cavity mass entrainment and modifying the local
simulate the spatial growth of the shear layer. The PSR residence time is
equivalence ratio and temperature distribution, which substantially
the mass in the reactor divided by the mass flow rate through the
affect the flammability limit with internal cavity fueling. Rasmussen
reactor. The flow residence time inside the cavity is estimated by
et al. [419] examined the stability of hydrocarbon-fueled flames in
assuming that the mass flow rate in and out of the cavity is dominated by
cavity flameholders, and large differences of lean blowout limit were
the compressible free shear layer. To approximate the convecting
noted between the cavity floor and cavity ramp injection schemes.
character of a mixing layer, flow enters the reactor from above and
Cavity ramp injection provided better performance near the lean
below the shear layer where no flow comes out of the reactor. The
blowout limit, whereas injection from the cavity floor resulted in more
density variation within the shear layer and the unsteady exchange of
stable flames near the rich limit. Flow and flame visualization revealed
mass at the cavity trailing edge was not considered in the approximate
that the cavity floor injection allowed both direct fueling into the shear
duration of flow residence time. The upper limit residence time repre­
layer as well as escape of substantial fuel through the shear layer, which
sents the minimum residence time where the mixture in the cavity can
resulted in localized fuel pockets and incomplete mixing, whereas in­
auto-ignite, and the lower limit residence time represents a minimum
jection from the ramp led to longer residence times and more uniform
residence required to sustain combustion within the cavity. This
mixing and combustion. In addition, Lin et al. [496] observed that the
simplified model indicates that the trace species diffusion should in­
fuel plumes injected from upstream injectors could decrease the rich
crease flame spreading rate and that heat loss reduce flameholding
blowout limit compared to the case with only direct cavity fueling. The
limits, which provides a design guideline for the size of cavity required
aerodynamic blockage from the upstream jet reduced air entrainment
to provide flameholding for a rectangular scramjet combustor. Details of
into the cavity. Choi [380] and Ghodke [498] simulated the cavity
the PSR analysis can be found in Ref. [500].
flameholding stability with cavity floor injection using LES. In stable
Unfortunately, the cavity flameholding experiments indicated that
combustion, a large vortical flow was formed in the aft region of the
modeling the cavity combustion as a PSR was physically inappropriate
cavity that provides an effective mechanism to transport the hot prod­
due to the high dependence of the lean and rich blowout limit on the fuel
ucts and to enhance mixing of the fuel-air mixture. In the near-blowout
injection scheme, that is, the passive/active injection into the cavity and
cases, small localized vortical flow structures were formed and seemed
the fuel injector location inside the cavity. The localized combustion
not to be as effective in mixing the hot products and transporting the
residing in the cavity shear layer and recirculation zone can be the
mixture.

52
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Changing freestream flow conditions could lead to cavity flame­ tunnel facilities.
holding blowout at various fuel injection configurations and fuel types. The Damkohler number (Da), a non-dimensional parameter, was
Donohue [493] investigated the effects of blowout temperature limit used to scale the theoretical and empirical relations for the flameholding
during dual-mode scramjet operation using external and internal fueling of premixed combustion in subsonic flows [503,504]. In these scaling
schemes and ethylene/JP-7 fuels. The blowout temperature was insen­ methods, the global parameters (e.g., equivalence ratio, reaction rate,
sitive to operating pressure with both ethylene and heated JP-7 fuels. ignition delay time, and the defined non-dimensional scaling parame­
Liquid fuel requires atomization that is sensitive to Reynolds number ters) are used to delineate the stability loop separating the region of
and operating pressure conditions. Lack of fuel atomization may result stable flames from the region of flame blowout. These globally defined
in reduced sensitivity to the freestream pressure. When a shock train is parameters can approximate the local stoichiometry condition in nearly
present, the higher pressure ratio produces lower velocities, higher homogeneous reactions encountered in the subsonic premixed com­
pressures, and higher static temperatures in and around the flame­ bustion devices. Nevertheless, in supersonic combustion systems like the
holder, making it easier to hold a flame at lower inlet air temperatures. scramjet engines, the majority of the fuel is consumed by the
Nevertheless, the blowout temperature was relatively insensitive to the non-premixed combustion and the progression of combustion is domi­
inlet Mach number at the isolator entrance, while the shock train loca­ nated by local stoichiometry conditions, which diminishes the effects of
tion and pressure ratio were nominally the same. Retaureau and Menon the globally defined parameters in premixed combustion. Hence, the
[497] conducted a sensitivity analysis on the stability domain of Damkohler number may need to be newly defined to consider the effect
self-sustained combustion with respect to the operating conditions. A of the local non-uniformity in the correlation of the non-premixed flame
small amount of hydrogen addition to methane led to a significant in­ blowout limits.
crease in stability, whereas ethylene addition had a less pronounced Driscoll and Rasmussen [106] developed a semi-empirical correla­
effect. The stability domain increased with the temperature of the su­ tion for flame blowout limits of non-premixed flames stabilized by shear
personic crossflow. layers located between the freestream and the recirculation zone in
Other parameters affecting the cavity blowout limit include the steps, cavities, and struts, in high-speed airflows. Fig. 67 shows the
backpressure, heat loss, and combustion vitiation effect. The back structure of a non-premixed flame stabilized by the shear layer when the
pressure can reduce the lean operational limit due to unfavorable fuel is injected directly into a wall cavity. The correlation assumes that
changes in air entrainment characteristics [496]. The majority of the the stabilized flame is composed by a lifted jet flame with a premixed
heat loss occurs at the wall and the shear layer, which can potentially be flame base, and propagates along the stoichiometric contour into the
an important parameter in blowout [497]. Experimental studies showed locally premixed mixture within the shear layer. This propagation speed
no distinguishable difference between the clean-air and vitiated tests of depends on the unburned gas temperature just upstream of the flame
flameholding in a cavity flameholder [499]. It was suggested that the base, which is elevated by the heat transfer from the cavity recirculation
compensation for vitiation effects on flameholding limits in directly zone. Flame blowout occurs when the upstream propagation speed of
fueled cavity flameholders was unnecessary in combustion heated wind the flame base cannot be matched by the gas velocity along the stoi­
chiometric contour. Rich blowout occurs when the flame liftoff distance
moves downstream approaching the length of the recirculation zone. On
the other hand, lean blowout limit occurs when the length of the flam­
mable zone along the stoichiometric contour reduces and the flame base
moves upstream until the flame propagation velocity is less than the
oncoming gas velocity. This correlation analysis leads to a definition of
the local Damkohler number based on the semi-empirical expression of
the characteristic time scales relevant to a non-premixed flame. A
detailed expression of the Damkohler number which correlates the flame
blowout limits measured for non-premixed flames can be found in
Ref. [106]. This analysis avoids the assumption that the residence time
of reactants in a well-stirred homogeneous reaction zone is matched to a

Fig. 67. Shear layer stabilized nonpremixed flames in a Mach 2.0 supersonic Fig. 68. Flame hydroxyl (OH) chemiluminescence of an annular cavity
cavity flameholder [106]. flameholder [506].

53
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

global chemical reaction time. Although the modified Damkohler 2.3.7. Role of cavity flameholder
number could adequately correlate the high-speed non-premixed flame Although the cavity flameholding phenomenon has been extensively
data, it was shown that additional research was needed to improve the studied in various configurations of flowpath geometry, cavity shape,
correlation by quantifying entrainment rate, temperature and species and fueling scheme, a general set of rules to design optimized cavities is
concentrations in the recirculation zone, and the flowpath taken by the still undefined. The main reason is the extremely wide range of condi­
fuel from the injector to the shear layer. tions over which a scramjet combustor is expected to operate. At low
Unsteady flame propagation in corner flow regions upstream of speed, with the scramjet operating in ramjet mode, combustion occurs
cavity flameholders needs to be accounted for in the cavity flameholding subsonically over a short distance downstream of the fuel injectors. In
limit in rectangular flowpaths. In rectangular flow channels, the corner this case, it was observed that a cavity flameholder placed at the end of
boundary layer effects can introduce unsteady flow distortions and the heat addition process can promote the formation of a thermal throat
unique changes to the flame dynamics. Donohue [493] observed sub­ and the re-acceleration of the flow from subsonic to supersonic. Under
stantial unsteady combustion both inside the cavity and when it prop­ these conditions, no heat release occurs in the diverging part of the
agated upstream in a direct-connect dual-mode scramjet combustor. The combustor that acts merely as an extension of the nozzle. At take-over
significant unsteady combustion occurring upstream of the flameholders Mach numbers, the flow in the combustor is supersonic but the tem­
should be accounted for in the flameholding limit since it can modify the perature is insufficient for auto-ignition. The cavity, in this case, has the
effective aerodynamic size of the cavity flameholder. Yentsch and Gai­ fundamental role of providing a suitable ignition environment because
tonde [135–137,481] found that the boundary-layer separation from of the low speed and high temperature. It is demonstrated that with a
corner flow effects dominated the combustion mode transition in small fraction of the fuel injected and ignited inside the cavity, a stable
dual-mode scramjet operations. These corner flow effects can be mini­ pilot flame can be established and used to ignite and promote combus­
mized in axisymmetric and/or elliptical flowpaths. The flame stability tion in the bulk supersonic flow. At higher flight speeds when auto­
limit of cavity flameholders in rectangular combustors cannot be ignition of the bulk flow becomes possible, direct injection of the fuel
directly extrapolated to that in the circular combustors. In rectangular into the cavity is no longer necessary. The flame stabilization mecha­
combustors, the cavity flameholders exchange mass flow with the nism in cavities without direct fuel injection appears less obvious. The
freestream through unsteady shear layer motion and corner flows. The literature reviewed does not show clear evidence that under these con­
cavity in the circular flowpath entrains the freestream into the recircu­ ditions a sustained stabilized flame can be established inside the cavity.
lation region through the shear layer only. The cavity flame blowout A plausible explanation can be found in the inherent difficulty of
limit depends on the cavity mass flow exchange rate, hence, the cavity guaranteeing the necessary conditions for ignition (both composition
flammability limit in the rectangular combustor does not necessarily and temperature), by relying uniquely on the flow entrainment prop­
apply to the circular combustor. Liu et al. [505] investigated cavity erties of the cavity. It seems therefore unlikely that with unfueled cav­
flameholding in an axisymmetric scramjet in a high-enthalpy Mach 4.5 ities the chemical contribution to flame stabilization might be
freestream. The axisymmetric scramjet emphasizes generic combustion significant, and the flame intensification observed experimentally is
and flameholding dynamics without the corner boundary-layer effects expected to be mainly the effect of enhanced mixing provided by the
that can distort flame propagation in rectangular geometries. Fig. 68 recirculating flow.
shows the flame distribution around the annular cavity flameholder An area of great importance to cavity flameholding design is to
[506]. There existed no apparent flame stabilized in the cavity volume maximize the thrust generation. Billig [507] defined an optimal heat
with upstream fueling though the cavity size was sufficiently large and addition process in a scramjet combustor, i.e., the minimal entropy rise
total pressure loss was little. The flammability limit in the axisymmetric for a given stagnation temperature rise and a set of initial conditions,
flowpath needs to be further examined to identify the dominant which can be realized when the flow reaches the sonic condition at the
impacting factor from the mass exchange rate, strain rate of flame, etc. combustor exit. Heat addition at the incorrect Mach, which does not
achieve this effect, could result in major pressure losses in the supersonic

Fig. 69. (a) Normalized pressure distributions and (b) OH-PLIF images (ϕ = 0.81) of axisymmetric scramjets with a cavity (left) and without a cavity (right) [508].

54
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

combustion process. During optimal heat addition, the effect of heat the thermal choking cannot be achieved due to the excessive area relief.
addition rate overtakes that from the combustor area relief, such that the Fotia and Driscoll [479] also observed that the pressure decreased
Mach number of the supersonic core decreases until the balance of heat monotonocally along the rectangular flowpath after the cavity in both
addition and area relief is reached at the combustor exit, resulting in the scram- and ram-mode operations of a direct-connect 2D rectangular
near sonic conditions. combustor. The flow was choked at the cavity ramp before entering the
Liu et al. [508] performed comparative experiments using axisym­ diverging section during the scram-to-ram mode transition. For both of
metric scramjet models with and without a cavity flameholder followed the two cases, the pronounced flame stabilized in the cavity during
by a 2◦ diverging section. It was observed that thermal choking could be scram-mode operation significantly modified the flame distribution due
achieved with the integration of a cavity flameholder, while no balance to the agglomerated heat release mainly in the cavity flameholder. As
of heat addition and area relief could be reached due to insufficient heat such, the diverging combustor serves merely as an extension of the
release when the cavity was replaced by a constant-area duct. Fig. 69 expansion nozzle downstream which should be avoided in a full scram
shows a scram-mode operation close to the optimal heat addition pro­ mode. When the dual-mode scramjet operates in the ram-mode, signif­
cess [508]. In this configuration, the combustion is successfully stabi­ icant combustion occurs inside the cavity volume and the diverging
lized in the diverging section with the aid of a cavity. Thermal choking combustor’s role is diminished, as observed in Ref. [506]. In ramjet
occurs at the pressure inflection point when the Mach number reaches mode, a long diverging combustor is pejorative for performance, but
sonic conditions. Downstream of the inflection point, the area expansion that is the price to pay to have an engine able to work over many
dominates the aerodynamic process and combustor serves as an different regimes.
expansion nozzle. In this type of geometry, it was observed that the
cavity enhances the combustion by mixing enhancement rather than the
continuous radical supply due to a lack of pronounced flame inside the 2.4. Energetic enhancement
cavity volume. Optimizing this balance between the heat addition and
area relief is one of the key optimization targets of combustor design. Plasma-assisted combustion is an energetic technique to enhance the
Insufficient area relief can lead to unexpected flow choking at an performance of combustion systems by coupling either a non-
early fueling stage [509], while excessive area relief can bring about equilibrium or thermal plasma discharge with combustion reactions to
abrupt flow expansion leading to flame quenching and combustion ef­ impact the chemical reaction rates. Using plasmas for ignition and
ficiency deterioration. Liu et al. [506] investigated the effect of the combustion enhancement dates back many decades, however since the
combustor divergence on cavity flameholding phenomenon and early 2000s, the field has seen a resurgence and has become a major field
dual-mode transition in an axisymmetric flowpath by comparing the 2◦ of study for many research groups around the world. Previous studies in
conical angle with a 5◦ . Numerical 1-D analysis showed that the diver­ plasma enhanced combustion have examined: pre-treatment of the fuel
gence of the 5◦ could not be thermally chocked due to the excessive or reactants into hydrogen rich syngas prior to thermal oxidation [511,
divergence. When heat release was increased to induce choking, the 512], enhanced ignition of hydrocarbon fuels [513–520], increased
choking initiation point was in the diverging combustor for the 2◦ model stability of combustion at atmospheric pressure [111,521–525], changes
as shown in Fig. 69, which would be expected in an optimally balanced in chemical conversion efficiency [526], and enhanced combustion
geometry. However, in the 5◦ geometry, the choking point would not stability for supersonic combustion systems [104,115,527,528]. A wide
initiate in the diverging section and would abruptly occur in the cavity range of studies worldwide is discussed by Ju [116,529], Starikovskiy
when mass injection exceeded a certain point and aerodynamically [530,531], Sun [532,533], Starikovskaia [534,535], and Leonov [536].
blocked the flowpath, an unrealistic representation of actual flight Fundamental mechanisms of plasma interaction with flames include
conditions. thermal effect, chemical kinetics, and radical transport. A survey of the
Likewise, many other experimental results demonstrate intense plasma enhanced combustion literature reveals that plasma impacts
flames being stabilized inside the cavity. We reiterate that such types of hydrocarbon chemistry through: (1) localized ohmic heating which in­
flameholding can fail to maximize the thrust generation due to the creases the rates of reaction and transport, and in-situ fuel reforming, (2)
inappropriate heat distribution. Gruber and Smith [510] characterized radiation-induced electron excitation, and (3) rapid decomposition of
the axisymmetric scramjet combustor flowpath in a direct-connect the fuel into smaller hydrocarbon molecules by chemically active spe­
combustor facility. The wall pressure peaked at the cavity ramp and cies. When a plasma is formed, the gas becomes conductive due to
decreased monotonically in the diverging combustor for all fueling rates. release of free electrons. Electromagnetic energy can be deposited into
It indicated that the flow was always choked at the cavity ramp before the flowfield by Joule heating, which can increase the local temperature
entering the diverging section. The flow Mach number increased of the gas and accelerate the combustion reaction rates. This is the case
throughout the diverging combustor. In this experimental configuration, for the thermal arc produced by a spark plug which generates a kernel of
high-temperature gas in the mixture allowing the chain reaction process

Fig. 70. (a) Classification of different type of plasmas. (b) Mechanisms of interaction between plasma and flame. From Ref. [116].

55
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

to start. Gliding arcs and ns-discharges operates in the non-equilibrium role of radicals [581]. Addition of hydrogen (30% in volume) to the
regime and are characterized by electron temperatures, Te = 1–10 eV, nitrogen plasma jet was found to give the best performance, being able
much higher than the kinetic temperature of the gas. These discharges to ignite all the fuels tested at room temperature with the minimum
produce limited thermal effects but seed the flow with highly reactive power input (2.5 kW). This effect was explained by the increased plasma
particles such as ions, radicals, and excited species. Chemical kinetics temperature due to combustion of the hydrogen with the air crossflow
can be modified by chain branching effect and reduction of the ignition [581]. On the other hand, ignition of hydrocarbon fuels was obtained
delay time. Large gradients and self-induced magnetic forces can make a only with a pure O2 jet because of the superior kinetic enhancement
plasma inherently unstable and dominated by turbulent transport [537]. effect of the O radical compared to N [583].
This characteristic can be exploited to generate turbulence and enhance In the case of downstream fuel injection, the lifetime of the radicals
mixing in a non-premixed reacting flow such as those occurring in a was found to play an important role. Despite the O radical being more
scramjet engine. Fig. 70 shows the mechanisms of combustion reactive, its short lifetime makes it less than optimal for combustion
enhancement by plasmas in different regions in terms of electron tem­ enhancement. A later study [582] showed that, in these conditions,
perature and number density [116]. N2–O2 mixtures are more effective than pure oxygen or pure nitrogen
In supersonic/hypersonic flows, the plasma-based combustion and feedstocks for ignition of both hydrogen and hydrocarbons. These effects
flow control techniques are used in both reacting and non-reacting were attributed to the formation of NO and NO2 radicals that have
flows. Plasma-assisted combustion has been proved effective to extend longer lifetimes compared to O and provide a catalytic effect on the
the flammability limits and ignition reliability in flameholding devices low-temperature chain-branching reaction of hydrocarbon fuels.
(cavities, steps) and jet-crossflow configurations [536]. Plasma-actuated Nevertheless, Abe et al. [584] observed that NOx radicals were signifi­
flow control in supersonic/hypersonic non-reacting flows has also been cantly less effective in igniting heavier hydrocarbons like propane
extensively investigated to extend the vehicle operationality over a wide (C3H8). Abe et al. [584] and Takita et al. [582] also investigated the
range of the flight envelope. These studies include applications such as possibility of further enhancing the ignition capabilities of a plasma
the boundary layer transition [538–540], shock generation and miti­ torch by supplying a small quantity of HO2 in addition to NOx. Nu­
gation [108,541–543], mixing enhancement [115,544–546], unstart merical simulation showed that HO2 is able to enhance the catalytic
control [547–549], inlet effective geometry control [550], etc. Addi­ effect of NOx at lower pressures [585].
tionally, utilization of magnetohydrodynamic (MHD) conversion of Several studies have addressed the influence of the relative geometry
momentum and energy in weakly ionized plasmas at hypervelocity between fuel injector and plasma torch [576,586–588]. Wagner et al.
provides advanced aerospace applications, such as heat flux mitigation [576] proposed an experimental configuration with several key features:
[551–560] and flight control [561] of re-entry vehicles, control of hy­ pilot fuel injectors, a step providing flow recirculation, a plasma-torch
personic boundary layer separation [562–566], and energy extraction located in the recirculation region behind the step (in order to create a
for directed-energy weapons in MHD-energy-bypass scramjets premixed combustion zone), and primary fuel injectors located down­
[567–573]. Nevertheless, the application of plasmas in supersonic flows stream. They found that the recirculation zone was able to sustain and
remains a challenge due to the limited range of action and low energy stabilize the flame once the torch was extinguished. Sato et al. [578]
efficiency particularly in electrically-driven plasmas [536]. In this sec­ tested three different injection schemes: a single orifice on a side wall, an
tion, the plasma-assisted combustion methods available for combustion array of four orifices on a side wall, and a total of nine orifices on both
stabilization in hypersonic flows are summarized, which include mainly side walls, with the walls separated by 38.4 mm. All the injectors were
plasma torches, laser-induced plasmas, and non-thermal plasmas. located 12.8 mm downstream of a 3.2 mm step that provided flow
Non-reactive plasma-based flow control will not be dealt with further in recirculation. The plasma torch was located on a side wall 20 mm up­
this review. stream of the step. It was found that the plasma torch was able to
effectively ignite the fuel jets on the same side. Ignition started from the
2.4.1. Plasma torches orifice directly downstream of the torch and then propagated to the
Early plasma assisted ignition experiments for scramjet propulsion others. Whether the flame persisted or extinguished with the plasma off
were conducted in the 1980s using plasma torch devices to ignite a su­ depended on the combination of flow stagnation temperature and
personic (Mach 2) air-hydrogen mixture [574–578]. Wagner et al. [576] equivalence ratio. Ignition of jets on both sides was later obtained by
found that a 1 kW-plasma torch, supplied with a 1:1 volumetric mixture modifying the top wall to include pilot fuel injectors, an additional
of argon and hydrogen and placed in the recirculation region of a plasma-torch igniter and a cavity recirculation region [587].
rearward-facing step, was able to drastically reduce the minimum Shuzenji et al. [589] considered a configuration with two plasma jets
stagnation temperature for ignition from the 1470 K required for in tandem and investigated the effects of their separation and relative
auto-ignition to room temperature (290 K). They also found that the mass flow rates. The plasma torches were located downstream of a single
kinetic effect produced by the hydrogen atoms had a predominant role fuel injector and upstream of a recirculation step. They observed that a
over thermal effects. In fact, when running the plasma torch with pure two-stage plasma torch with a separation of 20 mm and equal flow rates
argon, ignition occurred only at total temperatures above 890 K. This was able to outperform a single-stage in igniting methane fuel in a Mach
observation agrees with the calculations performed by Leonov et al. 2 airflow, requiring 4400 W instead of the 5500 W required by the
[579], where they compared the ignition of a stoichiometric air-H2 single-stage. Matsubara et al. [590] were able to reduce the
mixture at Mach 2.5 induced by a glow discharge and by simple heating plasma-torch power required to ignite hydrogen in a Mach 2 airflow
with the same total energy deposition. The results show that, with pure from 3.8 kW to 2.4 kW by installing a 10 W dielectric barrier discharge
thermal ignition, the induction length is 3.5 times greater than with (DBD) upstream of the plasma torch. Chemical kinetics simulations and
plasma ignition. spectroscopic measurements indicated the O3 radicals supplied by the
The effects of different composition of feed stock gases were inves­ DBD discharge as the probable reason for the reduced ignition delay
tigated experimentally and numerically by Takita et al. [580–582], time of stoichiometric H2/air.
simulating the ignition of hydrogen and hydrocarbon fuel jets in a su­ Bonanos et al. [591,592] studied a configuration in which fuel was
personic air crossflow (Mach 1.8) by means of a plasma jet. They found injected through four vectored jets arranged to form an aero ramp [338],
that the composition effects are drastically different depending on the with a plasma torch igniter located downstream. This configuration was
position of the fuel jet with respect to the plasma jet and the type of fuel. successfully applied to both hydrogen and ethylene fuel over a wide
With upstream fuel injection of hydrogen, pure O2 and pure N2 plasma range of equivalence ratios and stagnation temperatures. The
jets were found to have comparable ignition performance, and both counter-rotating vortices generated by the aero-ramp provided excellent
were more effective than a pure Ar jet, indicating again the fundamental mixing with the advantages of greater controllability and reduced

56
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 71. Sequence of flame development as a consequence of laser induced plasma [596].

blockage compared to a physical ramp. The optimal injector-igniter vectored jet and aerated-liquid normal jet, and two different torches: a
distance that maximizes the combustion efficiency was found to be DC-torch and an AC-torch operated with air or N2 feedstock. Chem­
dependent on the type of fuel and equivalence ratio. Jacobsen et al. iluminescence imaging of the flame plume, NO- and OH-PLIF were used
[593] further investigated torch/fuel-injector interaction for both as the main diagnostic tools. For similar power inputs, the DC-torch
gaseous (ethylene) and liquid (JP-7) hydrocarbon fuels. They considered proved to be more effective in igniting the flame, whereas no signifi­
three different injectors: the aero-ramp injector of Ref. [338], a single cant effect of the feedstock composition was observed.

57
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

2.4.2. Laser-induced plasma [598] demonstrated that locating the LIP inside the fuel jet, in order to
Plasma generated by laser induced breakdown is becoming an locally dissociate and ionize the hydrogen molecules, is a better ignition
increasingly widespread option for a large range of applications, strategy compared to direct shear layer LIP ignition in terms of overall
including internal combustion engines [594], jet engines [413], gas OH production, provided that the hydrogen jet is diluted with 8% argon
turbines, etc. The advantages offered by laser ignition over other sources in order to extend the plasma lifetime. Lee et al. [599] investigated LIP
were first discussed by Ronney [595] and summarized as follows: 1) a ignition of an axisymmetric hydrogen slot-jet in a supersonic crossflow
greater spatial and temporal flexibility and control, 2) reduced thermal of non-preheated pure oxygen at Mach 3. A plasma TIK was generated by
losses due to absence of solid surface, 3) improved durability and 4) a 300 mJ laser pulse in the shear-layer at the edge of the slot-jet and
multi-point ignition capabilities. Ronney also proposed potential appli­ transported in the recirculation zone downstream where it was able to
cations to high-speed propulsion systems. successfully ignite a self-sustained flame. It was also observed that with
First evidence of successful laser ignition in a model scramjet was a free stream of 80% oxygen and 20% nitrogen the TIK was completely
provided by Brieschenk et al. [596,597]. The experiments were con­ quenched and no ignition occurred.
ducted in the T-ADFA free piston shock tunnel at a stagnation enthalpy Laser-induced breakdown ignition inside a cavity flame-holder was
of 2.7 MJ/kg and freestream velocity of 2075 m/s. The model scramjet investigated extensively at WPAFB [83,85,410,600–603]. Do and Carter
consisted of a 9◦ compression ramp and a constant area combustor. [84] demonstrated the possibility of performing simultaneous laser
Hydrogen fuel was injected in a transverse jet in a crossflow configu­ ignition and laser induced breakdown spectroscopy (LIBS). Using this
ration through an array of four holes located on the inlet ramp. The laser technique, it was possible not only to determine the ignition probability
beam was injected through a window located between the inner injector at a certain location of the flowfield, but also to obtain quantitative
holes and focused 8 mm above the surface, inside the shear-layer be­ information on local temperature and gas composition (in particular the
tween fuel jets and air crossflow. A Q-switched ruby laser with pulse fuel concentration) at the location probed. This was accomplished by
energy of 750 mJ was used to induce plasma breakdown. It was esti­ direct-spectral matching (DSM) [83]. A database of breakdown spectral
mated that about 54% of the laser energy was absorbed by the LIP. Using emissions was built using a variable pressure combustion chamber
schlieren and OH fluorescence, the authors documented the formation (VPCC) filled with an ethylene-air mixture over a wide range of mixture
and evolution of a combustion transient ignition kernel (TIK) from the density and fuel concentrations. By finding the spectrum in the database
breakdown point to the entrance of the isolator (Fig. 71) and suggested closest to the spectrum profile taken during an ignition test, it was
that a continuous stream of hydroxyl radical would require laser pulse possible to derive the gas density and atom concentration. The spectral
repetition at ~100 kHz. It was found that the early development of the window chosen for the experiments was 550 nm–830 nm, containing
TIK was scarcely influenced by the random turbulent fluctuations near multiple emission lines of N, O, C and H that exhibit high signal-to-noise
the injectors and exhibited a remarkable repeatability. Brieschenk et al. ratio. Using the DSM technique, McGann et al. [601] characterized the
cavity flow of the RC-19 continuous flow supersonic wind tunnel at
Mach 2 and a stagnation temperature of 590 K, using ethylene as fuel. A
total of 36 locations were probed with the LIP, determining those
causing ignition and sustained combustion. Fig. 72 provides a schematic
of the experimental setup and distributions of fuel mole fraction. It was
found that in order to obtain sustained combustion a fuel mole fraction
above 3.5% was necessary and these conditions occurred in the recir­
culation region near the front step of the cavity.
Extensive investigation of laser-induced plasma ignition in a cavity
flame-holder has also been recently conducted in China [408,604–606].
An et al. [606] investigated the effects of laser energy and breakdown
location on the ignition process using OH* (311 ± 5 nm) and CH* (430
± 5 nm) chemiluminescence visualization. The experiments were con­
ducted in a direct-connect vitiated-air facility at Mach 2.92, total tem­
perature 1650 K and total pressure 2.6 MPa, using ethylene fuel at a
constant equivalence ratio of 0.152. It was found that the middle part of
the cavity provides a higher ignition probability with a single pulse
(87%) probably due to the combined effects of mixing, initial kernel
strength, and low strain rate of the recirculating flow. The kernel
generated at this location is advected upstream towards the cavity front
step where it ignites the shear layer. The ignition time, defined as the
time necessary to reach a quasi-stable burning condition, was found to
depend on the laser pulse energy EP, 720 μs for EP = 303.6 mJ and 1120
μs for EP = 230.7 mJ. Similar tests were conducted by Li et al. [604] in a
dual cavity configuration with ethylene fuel injected upstream of the
cavities. The benefits of using a dual cavity arrangement are discussed in
Ref. [607]. Combustor entrance Mach number, stagnation temperature
and stagnation pressure (2.1, 947 K and 0.65 MPa respectively) corre­
sponded to flight conditions at Mach 4. Ignition was provided by a laser
pulse with energy of 960 mJ at 1064 nm, focused on the middle of the
upstream cavity. The dynamics of the ignition process, studied by
high-speed (40,000 fps) photography, was qualitatively very similar to
the experiments of An et al. [606], but in this case the time required by
the flame to fill the cavity was significantly longer, 1950 μs, despite the
higher pulse energy. This longer ignition time could be due to the lower
Fig. 72. (a) Experimental setup for the LIB ignition experiments [410]. (b) static temperature of the gas, 503.2 K instead of 609.9 K (computed with
Mapping of fuel concentration inside the cavity [601]. isentropic flow relations), or to the lower density and therefore lower

58
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

energy absorption during breakdown. Ignition of the second cavity recirculating region within the cavity. Furthermore, in the lean condi­
began at 2300 μs and was completed after 2950 μs, when quasi-steady tions tested (ϕ = 0.16), ignition was obtained only with the LIP in the
combustion was reached. Using a similar dual-cavity setup, Li et al. rear part of the cavity, whereas when the LIP was in the front part, the
[608] were able to successfully ignite liquid kerosene in a supersonic residence time was too short before the hot plasma convected into the
flow at Mach 2.52. Fuel was injected and atomized into the cavity up­ high speed shear layer. The authors also used LES numerical simulations
stream (C1) and ignited in the downstream cavity (C2) by a 183 mJ to study the mixing process prior to ignition and achieved a better un­
(65% deposited to the plasma) laser pulse. Successful ignition was derstanding of ignition. In particular, turbulent kinetic energy and scalar
determined by LIP location and overall equivalence ratio. Once ignited, dissipation rates were found to have a crucial effect on ignition. Cai et al.
the flame was able to propagate from cavity C2 to C1 establishing full [409] further characterized the ignition and flame stabilization pro­
sustainable combustion. Attempts to directly ignite the C1 cavity were cesses with various initial LIP locations with an identical ethylene
unsuccessful due to insufficient atomization and mixing. fueling scheme and identical freestream conditions. The transient igni­
The advantages of using two overlapped laser pulses for breakdown tion process depended on the initial LIP location. The initial plasma
was first discussed by Dumitrache et al. [609]. In synthesis, the first laser flame kernel generated in the foreside of the cavity may move down­
pulse acts as a pre-ionization source, lowering the breakdown threshold stream along the high-speed shear layer above the cavity. The initial
of the second. This technique was later applied to ignite atmospheric flame generated in the rear portion of the cavity may propagate up­
flames of hydrocarbon fuels [603,610]. Application to scramjet pro­ stream along the cavity floor and eventually to the shear layer where a
pulsion and supersonic combustion was studied by Yang et al. [605] in a stable flame could be established. Fig. 73 shows the typical LIP ignition
single cavity in a direct-connect combustor, which had a freestream process characterized using high-speed flame chemiluminescence and
Mach number of 2.52 and a total temperature of 1486 K. Room tem­ schlieren imaging. A two-regime ignition mechanism was proposed
perature gaseous ethylene fuel was injected sonically through a fuel based on the transient ignition phenomenon.
injector located upstream of the cavity front edge. It was shown that a 15 Hassan et al. [612] investigated the ignition and flame propagation
ns interval between two successive laser pulses provided enhanced in an ethylene-fueled cavity flameholder using LES. The airstream at the
ignition capabilities compared to a single pulse with the same overall combustor inlet was at Mach 2 with a total temperature of 589 K and a
energy. The authors also performed a sensitivity study with respect to total pressure of 483 kPa. Ethylene fuel was injected through a row of
pulse energy and time interval. For the single laser pulse ignition, it was evenly spaced injectors near the base of the cavity closeout ramp. The
found that higher laser energy increases the projected area of the igni­ unsteady behaviors of heat release were characterized by short-lived
tion kernel and shortens the ignition time. The maximum interval be­ ethynyloxy radical (HCCO) and the results were qualitatively
tween the pulses able to produce ignition enhancement was found to be compared with analogous experimental observations. The LIP kernel
50 μs Beyond this value, the effects of pre-ionization vanished, and initially spread to fill the back of the cavity in the spanwise direction,
ignition failed. Li et al. [611] performed similar dual-pulse laser ignition and then moved axially downstream to ignite the entire cavity. The
in a kerosene-fueled cavity in a Mach 2.92 airflow. There exists an observed unsteady flame propagation during the LIP ignition was caused
optimal time interval between the two successive laser pulses such that by the coupling between the ethylene fueling rate and the cavity pres­
the two induced flame kernels could merge together and propagate into sure rise due to heat release. At stationary conditions, the heat release
a stable combustion flame. was mainly stabilized in two distinct regions: in the shear layer and near
Cai et al. [408] performed laser ignition experiments in an the exit of the injector ports. Cai et al. [408] studied the LIP ignition
ethylene-fueled cavity flameholder at various fueling rates. The inflow using LES as well and demonstrated an increased probability of LIP
at the combustor inlet had a Mach number of 2.92, stagnation pressure ignition inside the subsonic zones and the rear side of the cavity with
of 2.6 MPa, and stagnation temperature of 1530 K. They described the ethylene fueling upstream of the cavity flameholder.
ignition process as being composed of four stages: plasma ignition,
plasma quenching, re-ignition, and stable flame. Stable combustion was 2.4.3. Non-thermal plasma effects
achieved only with the LIP located in the low speed (subsonic) or Non-thermal plasmas are characterized by a kinetic temperature

Fig. 73. Temporal evolution of CH* chemiluminescence (top) and schlieren imaging (bottom) of the LIP ignition process (ϕ = 0.3) [408].

59
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

ignition process. Fig. 74 shows the two configurations that were exam­
ined: one with the electrodes placed on the bottom of a cavity down­
stream of the injector [527] and one without a cavity and with the
electrodes placed between two fuel injectors [528].
In both cases, the plasma discharge effectively ignited a sustained
flame that was also maintained in plasma-off conditions. A significant
reduction of the ignition delay time was observed and explained with
the generation of reactive radicals by the plasma. A very similar
behavior was observed for hydrogen and ethylene fuels. Fig. 75 shows
the OH distributions superimposed on schlieren images, for the case of
Fig. 74b. A comparison was made among no-plasma, anode upstream of
the cathode, and cathode upstream of the anode. The last case was found
to be the most effective in terms of OH production.
A systematic study of the effects of inter-pulse coupling for high-
frequency nanosecond discharges was performed by Lefkowitz and
Ombrello [616] in a flowing stream (subsonic speed) of air and methane
at different equivalence ratios. Three regimes of pulse coupling were
identified depending on the inter-pulse time: namely fully-coupled,
partially-coupled and decoupled regimes. The fully-coupled regime oc­
curs when all the pulses are deposited into a small volume and has the
highest probability to develop into a freely propagating flame. The study
Fig. 74. Experimental setup for (a) the cavity ignition [527] and (b) ignition showed the existence of an optimal inter-pulse time able to provide both
between the jets [528]. high ignition probability and fast growth rate. This optimal value de­
pends on number of pulses, equivalence ratio, electrode distance, and
significantly lower than the vibrational and electronic temperatures. flow velocity. A similar study performed by Lovascio et al. [617] in a
Some examples of non-thermal plasmas are gliding arcs [111,613], constant volume combustion chamber using a propane-air mixture at a
nanosecond discharges, and dielectric barrier discharges (DBD). Since fixed equivalence ratio of 0.7 led to similar conclusions.
non-thermal plasmas do not significantly increase the local temperature Numerical simulations of high-frequency, plasma-assisted combus­
of the flow, they affect ignition primarily through the introduction of tion are extremely challenging due to the wide range of temporal scales
radicals and excited species into the mixture. These highly reactive involved that make these computations extremely expensive, especially
species allow the reaction process to bypass the chain initiating steps for 3D cases. Effective modeling of the fundamental physical phenom­
(typically characterized by high-activation energies) and therefore ena involved is crucial for obtaining accurate results. The choice of the
reduce the ignition delay time. This effect is also present in thermal model is dictated by the nature of the problem to be studied. Takana
plasmas, but it is overcome by purely thermal effects. Thermal and et al. [618] investigated nanosecond pulsed pin-to-pin discharges in 2D,
non-thermal effects of low-temperature plasmas are discussed in detail for a lean methane/air mixture at a pressure of 10 bar and a temperature
by Adamovich et al. [520]. of 600 K. A detailed kinetic model with 325 reactions and 53 chemical
There are several advantages of non-thermal discharges. Due to their species was considered. At an equivalence ratio of 0.5, it was found that
short duration, typically on the order of a few nanoseconds, they tend to more than 60% of the total electron energy was consumed by vibrational
be more stable than thermal plasmas (especially at high pressure), more excitation of the molecules and less than 5% was consumed by disso­
efficient and more controllable in terms of energy addition [515,516, ciation of oxygen. The total electron energy was found to be increasing
579,614,615]. In addition, they can be pulsed at very high repetition with the equivalence ratio. Wang [619] carried out a LES simulation in
rates, taking advantage of shot-shot interaction phenomena. The inter­ 3D of combustion enhancement by a repetitively pulsed plasma in a
ested reader is referred to the review by Starikovskaia [535] on the non-premixed turbulent air/methane flow. In this case, a three-species
subject. physics-based model was used to compute both body-forces and heat
Do et al. [527,528] experimentally demonstrated the capability of addition due to plasma-combustion coupling.
igniting both hydrogen and ethylene flames in a supersonic oxygen Ignition and flame stabilization in supersonic flows by transverse
stream. The experiments were conducted in an expansion tube at Mach short-pulse repetitive filamentary discharges were investigated exten­
numbers between 1.7 and 3. The nonequilibrium plasma was created by sively by Leonov and co-authors [104,614]. A variety of configurations
repetitive nanosecond (10–20 ns) pulses of 15 kV peak voltage at were tested, including a single filament across a circular duct [620,621],
50 kHz. Schlieren and OH-PLIF imaging were used to characterize the a flat wall with arrays of electrodes upstream and downstream of the fuel

Fig. 75. OH-PLIF and schlieren images superimposed for cases with (a) no-plasma, (b) anode upstream of the cathode, and (c) cathode upstream of the anode [528].

60
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 76. Development of turbulence in after-spark [545].

injectors [104,622,623], a cavity flame-holder and a backward wall step could be significantly increased for complete combustion reaction and
[104,105,579,614,620,624], using both hydrogen and ethylene fuel. sufficient thrust generation. Nevertheless, the combustor design needs to
Tunable diode laser absorption spectroscopy of the H2O molecule was be integrated to a scramjet-powered hypersonic system over a wide
used to quantitatively estimate the local temperature of the flow and range of flight envelope. At the high Mach number regime, combustion
water vapor concentration. These studies demonstrated the multi-step chamber density must be high enough to ensure that combustion takes
nature of plasma-assisted ignition. In the simplest case of two-stage place within a reasonable length such that the combustor length can be
ignition, the first stage consists of the production of active radicals compatible to the overall vehicle size. Increasing the inlet contraction
and fuel reforming by the plasma. Despite being characterized by bright ratio and the area ratio of thrust nozzle could increase combustion duct
emission, no significant heat release occurs and therefore this stage is pressure, reduce the combustor length needed for complete combustion,
named “cold combustion”. The second stage is when the temperature and generate more thrust. However, these benefits are marginal due to
and pressure rise occur and is therefore named “hot combustion”. the limitation on overall back pressure rise imposed by boundary layer
Experiments by Leonov and co-authors also showed [544–546,625] separation [80]. Reducing the combustor length and devising a com­
that transverse electrical discharges have an effect on mixing in super­ bustion mode different from the traditional diffusive combustion are
sonic flows. The schlieren images of Fig. 76 show the development of needed for integrated scramjet design over a wide range of flight Mach
turbulence between the electrodes in the after-spark phase. This numbers. Theoretical analysis of scramjet cycle performance demon­
behavior was explained [115] by a pinch instability on the plasma col­ strates significant benefits from premixed combustion, and the
umn with the formation of small density inhomogeneities along the finite-rate chemistry analysis suggests that premixing may be necessary
entire length of the discharge. These density gradients develop into for scramjet combustion in the flight Mach number range of 8–22 [122].
vortices by the Rayleigh–Taylor instability. In the case in which the Shock-induced combustion involving forebody fuel injection was
plasma channel has bends, an additional jet instability is present due to designed to achieve adequate mixing and combustion stability without
the interaction of converging and diverging shocks [626]. The turbu­ the use of complicated mixing-enhancement and/or flameholding de­
lence generated by this mechanism is expected to be applicable to the vices. Radical farming and oblique detonative wave are examples of
mixing enhancement of air/fuel in supersonic combustion. Nevertheless, shock-induced combustion mode proven to be effective in combustion
numerical simulations performed on a jet-in-crossflow configuration in
which a plasma is generated by a laser pulse [627], did not show sig­
nificant bulk mixing improvement due to the plasma.

3. Shock-induced combustion

With increasing hypersonic flight Mach number (M∞ > 8), sufficient
mixing and combustion cannot be achieved within the current
combustor length and at current pre-combustion pressures, which re­
sults in inadequate thrust generation. The overall length of combustor

Fig. 77. Schematic of radical farming mechanism in a scramjet [628]. Fig. 78. Regions of radical-farming behavior [628].

61
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

stabilization for hypersonic propulsion. 3.1.2. Concept validation


McGuire et al. [628] investigated numerically the dependency of the
3.1. Radical generation radical farming concept on the inflow conditions. The authors per­
formed 2D simulations of a scramjet engine combustor working with a
The concept of radical farming was developed to explain the exper­ fully premixed hydrogen-air flow. A parametric study on the stagnation
imental observations made in the T4 tunnel at the University of conditions of the flow revealed that for a given stagnation pressure,
Queensland, showing that combustion-induced pressure rise was initi­ radical farming behavior occurs only in a certain range of stagnation
ated at the location of the second shock reflection inside the combustor temperatures. Increasing the stagnation pressure results in this range
[118,628–634]. becoming narrower and shifting to a lower temperature, as shown in
Fig. 78. Within this range, heat release, temperature rise, and pressure
3.1.1. Combustion stabilization mechanism rise are observed to start at the second or successive hot pocket, indi­
In radical-farming configuration, a mild inlet compression system cating that the first pocket is acting as a radical farm. Below the range,
was used to alleviate excessive total pressure loss from shock generation. the temperature in the first hot pocket is insufficient to produce enough
Though milder freestream compressions lead to lower pressures and radical species; no ignition occurs, and therefore the flow is essentially
temperatures within the combustor section, the associated shock inter­ non combusting. Above the range, the temperature is sufficient to ignite
action with the developing viscous layers produce recirculation regions combustion directly in the first hot pocket.
amenable for combustion reactions. Odam and Paull [629] investigated The results predicted by radical farming theory were later verified
this phenomenon using the experimental setup shown in Fig. 77, where experimentally by the same authors [630] using a planar geometry
the combustor flow consists of a series of oblique shocks and expansion scramjet combustor. The experiments were carried out in a shock tunnel
fans reflected from the walls. Fuel was injected at the inlet to allow more at stagnation conditions corresponding to Mach 8.7–11.8 flight. Fig. 79
time for mixing. The diamond-shaped regions bounded upstream by shows the model geometry, wall pressure measurements, and OH-PLIF
intersecting shocks and downstream by intersecting expansions were images taken at the second region of the shock/expansion structures.
characterized by high temperature and pressure where the chemical The pressure measurements supported the radical farming behavior,
reactions become active. Although the residence time of the flow in with combustion induced pressure rise starting from the second hot
these regions is too short to achieve complete combustion or even pocket. The OH-PLIF images confirmed the presence of OH radicals
ignition, radical species are formed and seeded into the flow. For this ahead of the ignition point, in the region between the first two hot
reason, these regions are called radical farms. Outside of radical farms, pockets on the lower side. No activity was observed in the upper side hot
the radicals freeze due to the lower temperature, but once the temper­ pockets due to poor air-fuel mixing in this region. Despite not being fully
ature rises again in the successive farms, the radicals re-activate accel­ conclusive, these results support the radical farming hypothesis. The
erating the ignition process and leading to sustained combustion. The authors [631] also directly visualized the OH chemiluminescence in the
authors point out that the performance of a scramjet based on radical first pocket initiating the ignition process and accelerating the com­
farming was not significantly affected by the inlet contraction ratio as bustion process in the next downstream hot pocket in a 2D inlet-fueled
was the case in regular scramjets. The main parameter to consider in this scramjet configuration at stagnation conditions corresponding to Mach
case is the combustor height-to-length ratio, which determines the shock 9.7 flight. Finite-rate chemical kinetic analysis revealed that the com­
structure. plex interdependence of different chemical reaction groups within the

Fig. 79. (a) Experimental setup, (b) pressure distribution at different times, (c) OH-PLIF visualization [630].

62
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 80. Effects of inlet injection through a porous wall and a porthole [633,634].

hydrogen-oxygen reaction system led to a chain-branching explosion. [644–646] were performed. Hydrogen fuel was injected in the down­
The explosion governed the radical farming process in the combustion stream direction in a heated supersonic airstream and then ignited after
flowfield with low mean temperature and pressure. passing through a normal shock, which was generated by an under­
Efficient mixing is a key element for the performance of a radical expanded jet or formed by the intersection of two oblique shocks. Later
farming scramjet. A porthole injector at the inlet ramp was initially it was found that the combustion after a normal shock in such config­
considered for the detonation-driven scramjet and later for mixing urations would not be classified as detonation because the normal shock
enhancement and combustor length reduction in conventional diffusive was formed by an arrangement of wedges and shockwaves rather than
scramjets. Effects of the inlet porthole injection on mixing enhancement, the heat release from the downstream combustion process. The com­
overall combustion distribution, and cavity flameholding were investi­ bustion phenomena in these experiments were initiated by the sudden
gated in a REST scramjet [143,635–639] and a rectangular 2D scramjet temperature rise in the airstream created by a normal shock, in which
[640–642]. For radical farming type scramjets, Capra et al. [632,634] the static temperature exceeded the fuel ignition temperature. Most
investigated numerically a fuel injection scheme through a porous sur­ importantly, the pressure and temperature produced by a normal shock
face located at the inlet. The study showed that despite a lower pene­ at hypersonic Mach numbers result in high structural loads, extreme
tration depth, the porous wall generally provides better performance cooling problems, and a high degree of chemical dissociation that will
compared to a conventional port hole. With porous wall injection, delay the completion of the exothermic reactions for thrust production
mixing was improved by the larger interaction surface, reducing ignition [647]. Normal shock-induced combustion was thus undesirable, and the
delay, and increasing combustion efficiency. The porous wall injection oblique shock-induced combustion gained greater interest.
suffered a reduced pressure loss because the attached oblique The concept of combustion phenomenon induced by an oblique
fuel-injection shock was weaker than the shock generated by a jet in a shock was initially demonstrated in a series of investigations mimicking
crossflow, as shown in Fig. 80, and the associated stronger a scramjet inlet with two standing oblique shocks at low Mach numbers
shock-expansion train ingested by the engine enhances spreading of fuel of 1.6–1.7 [648–651]. Due to the ignition delay, a separation between
and promotes rapid growing of the mixing layer. The feasibility of shock and exothermic reaction zone may exist in the shock-induced
porous wall injection was demonstrated experimentally by Capra et al. combustion [122]. The shock shape and strength can be affected by
[633] using a ceramic matrix composite (CMC) insert as a porous close-coupled exothermic reactions. The term ‘oblique detonation wave’
injector. The experiments were performed at a stagnation pressure and (ODW) applies for a close-coupled oblique shock, in which the shock is
enthalpy of 40 MPa and 4.3 MJ/kg, respectively. This study also inves­ driven by a downstream pressure rise from the exothermic reactions.
tigated the effects of oxygen enrichment of the hydrogen fuel. Increases ODW engines include the shock-induced combustion ramjet and the ram
of oxygen concentration by 13%, 15% and 17% were shown to signifi­ accelerator. Comparisons between a conventional diffusive-burning
cantly enhance the combustion performance of a radical farming scramjet engine and ODW engines are shown in Fig. 81. ODW com­
scramjet in terms of ignition delay and overall combustion length. bustion can be stabilized by either a 2D planar wedge or a 3D body such
as a cone-nosed or blunt-nosed hypersonic projectile.
3.2. Oblique detonation wave
3.2.1. Mechanism of shock-induced combustion
The applicability of a standing detonation wave to a reaction engine The shock-induced combustion in the premixed gases is usually
was originally explored by Dunlap et al. in 1958 [120,121,643]. Their described using the Rankine-Hugoniot (R–H) relations in the pressure-
theoretical analyses suggested that detonative combustion was able to specific volume plane as shown in Fig. 82. The R–H relations describe
significantly extend the operational range of a ramjet engine to higher the relationship between the states on both sides of a shock wave or a
Mach numbers up to 22, maintaining comparable performance. A series combustion wave (deflagration or detonation) in a one-dimensional
of experimental studies on normal shock-induced combustion flow in fluids or a one-dimensional deformation in solids. The R–H

63
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 81. Schematics of a conventional scramjet and oblique detonation wave engines [652,653].

conditions consist of the conservation of mass, momentum, and energy


of gas flow having a discontinuity. The Rayleigh line delineates con­
servation of mass and momentum,
p2 − p1
= − m2 , (33)
1/ρ2 − 1/ρ1

where p is pressure, ρ is mass density, m is the mass flow rate per unit
area, subscript represents the conditions upstream and downstream of
the wave. The Hugoniot equation combines the conservation of mass
and momentum and the conservation of energy,
( )
1 1 1
h2 − h1 = + (p2 − p1 ), (34)
2 ρ2 ρ2

where h is the specific enthalpy of flow. The change in enthalpies rep­


resents the combustion heat addition in an adiabatic reacting flow. In
Fig. 82. The Rankine-Hugoniot relation in the pressure-specific volume plane conjunction with the relation between the upstream and downstream
showing the sections of the curve corresponding to the various combustion equation of state, the intersection of above two equations in the p-1/ρ
domains [654]. plane determines the downstream conditions if the upstream conditions
are given. Detailed explanation can be found in the corresponding
chapter of ref. [654].
The Chapman-Jouguet (C-J) condition states that waves propagate at

64
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

a velocity at which the reacting gases just reach sonic velocity (in the operation. Rubins [651] investigated the effects of hydrogen injection in
frame of the leading shockwave) as the reaction ceases (point U in the inlet of a two-shock inlet hypersonic ramjet using a simplified model
Fig. 82), i.e., detonation. In the system of the detonation wave with a at Mach 9–22. It was found that the low molecular hydrogen injection
leading shock, gases enter at supersonic velocity and are compressed reduced friction and improved performance. Most importantly, it was
through the shock to a high-density, subsonic flow. This sudden change shown that reaction kinetic rates rather than mixing factors control
in pressure initiates the chemical energy release. The energy release re- combustor length at higher altitudes and Mach numbers. Nevertheless,
accelerates the flow back to the local speed of sound. The sonic plane methods to premix fuel and freestream were not suggested. Sislian et al.
forms a choke point that enables the lead shock, and reaction zone, to [664–670] numerically explored cantilevered ramp injectors for fuel
travel at a constant velocity without disturbance from the expansion of delivery in high enthalpy, high Mach number airflow corresponding to
gases in the rarefaction region beyond the C-J plane. Theoretical anal­ shcramjet flight conditions. These studies were limited to hydrogen/air
ysis shows that the upper C-J point (U) is detonation driven by com­ mixing in an external-compression inlet at a flight Mach number of 11
bustion heat release while the lower C-J point (L) is deflagration driven and a dynamic pressure of 0.67 bar, with use of an array of cantilevered
by gas expansion. Usually the real combustion detonations have com­ ramp injectors. The cantilevered ramp injectors were particularly suited
plex 3D structures with parts of the wave traveling faster and others to keep the combustible mixture away from the boundary layers of the
slower than average, which poses combustion instability problems in duct walls in shcramjet engines. The injector array spacing, the single
establishing a stable detonation reaction with a normal or oblique and alternating injection angles, and the sweeping angle were varied to
leading shockwave. maximize the overall mixing efficiency while minimizing the total
pressure loss. It was shown that the mixing efficiency was increased by
3.2.2. Development of ODW engines 31% for a convective Mach number of 0–1.5, while the stagnation
The class of oblique detonation wave ramjet was investigated as a temperature was reduced by 10%. The fuel injection in the inlet was
part of studies on external burning in supersonic streams at APL, Johns found to increase the thrust potential with a gain exceeding the losses by
Hopkins University [655–657] in the 1970s. It was proved that an 40–120%, while skin friction in the inlet was estimated to make up
external ramjet, i.e., burning beneath a flat-topped, tri-angular airfoil, 50–70% of the thrust potential losses. The effects of incomplete fuel/air
could produce net thrust as well as lift at Mach 5. But the specific im­ mixing on the propulsive performance of a mixed-compression inlet
pulse decreased drastically when the external burning device was used shcramjet at different flight Mach numbers are shown in Fig. 83 [663,
from drag reduction to thrust production. Billig [117] considered 671].
external burning impractical as an accelerator; the external burning Operation of ODW engines requires premixing the fuel and air while
configuration must be used in tandem with a primary thrust-producing avoiding ignition upstream of a stabilizing wedge. Veraar and Mayer
device in plausible applications. In the external burner, the pressure and [672] performed a proof-of-principle experiment using an axisymmetric
temperature are lower, the residence time for completion of heat is dual cone test to demonstrate the possibility of injecting hydrogen into a
shorter, and the drag penalties from flame stabilization devices are much high enthalpy supersonic air flow without causing premature ignition
greater due to the high dynamic pressure. The internally ducted engine and subsequently inducing combustion of the mixture by a strong obli­
was considered more practical for shock-induced supersonic combustion que shock. The supersonic airstream had a speed of Mach 3.25 and a
designs. Morrison [658,659] reported the propulsive performance of stagnation temperature of 1450 K. A thermocouple installed on the cone
external and mixed compression ODW ramjets based on simplified surface, regular video images, and shadowgraphs were used to charac­
one-dimensional analysis. It was concluded that the internally ducted terize the shock-induced combustion phenomenon. Fig. 84 shows the
ODW ramjet offered a great potential to extend the useful range of schematic of the test geometry and video images of test cases with and
ramjet flight Mach number to 6–16 and above. Menees et al. [660–662] without hydrogen injection. In conjunction with the temperature his­
numerically validated the concept of a trans-atmospheric vehicle pow­ tories recorded on the top and bottom cones, it was confirmed that the
ered by an ODW engine and provided experimental proof-of-concept shock-induced combustion occurred in both the shock-induced bound­
studies on an ODW engine using an in-stream strut fuel injector in an ary layer separation zone and the inviscid flow field outside of the
arc-heated hypersonic wind tunnel. boundary layer. The results also showed that a small deviation in the
Injecting and mixing the fuel in the inlet diffuser is critical to incline angle of the test object with respect to the air flow could result in
decouple the fuel/air mixing and burning process of shcramjet engine hydrogen being captured and subsequently burned in the boundary
layer separation zone resulting in extremely high local heat loads.
A major advantage of an ODW engine is its smaller size in compar­
ison to a conventional diffusive scramjet. Ashford and Emanuel [673]
analyzed the performance of an ODW engine and a diffusive scramjet
engine with either constant pressure or constant area combustion for
hydrogen and methane fuel. The performances were evaluated at Mach
8–20 and altitude 12–36 km. The ODW engine had comparable engine
performance to diffusive scramjet engines with a difference in specific
impulse of 15%. Couture et al. [674] compared the performance pa­
rameters between scramjet and shcramjet at Mach 7–12 in terms of
intake performance, combustor performance, and lift-to-drag ratio. It
was found that the shcramjet range was slightly better than the scramjet
for combustor efficiency up to only 0.8. This reduction in specific im­
pulse can be offset by the benefits from the reduced size. O’Neill and
Lewis [675] suggested a 50% reduction in the length of the engine at
Mach 10 cruise configuration when an ODW engine was used in
waverider-based hypersonic vehicles. Chan et al. [676] compared the
propulsion performance of a scramjet and shcramjet at Mach 11 and an
altitude of 34.5 km. Hydrogen fuel was used in the numerical study. To
evaluate the benefit of ODW propulsion, the scramjet and shcramjet
Fig. 83. Hypersonic engine specific impulse as a function of flight Mach engines shared the inlet configuration, fuel injection system, fuel/air
number [663]. equivalence ratio, mixing/combustor duct, and methodology of nozzle

65
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Fig. 84. Proof-of-principle experiment of a shock-induced combustion ramjet using hydrogen fuel [672].

Mach numbers. A high equivalence ratio limits the maximum temper­


ature in the combustion zone and thereby reduces the fraction of
dissociated species in the nozzle expansion process. For instance, typical
values of overall equivalence ratio ϕ = 1.3 at flight M∞ = 12,
ϕ = 4 at M∞ = 20, and ϕ = 6 at M∞ = 25, were recommended as impor­
tant design guidelines for high Mach number flight.
The ram accelerator, which is a subset of hypervelocity accelerators,
is capable of launching projectiles at velocities of 8 km/s or above [677].
In experiments, a stationary tube is filled with a premixed gaseous
propellant and is closed at each end by thin diaphragms. A cone-nosed or
blunt-nosed hypersonic projectile having a diameter smaller than the
bore of the accelerator tube is accelerated from rest by a conventional
gas gun or light gas gun. The projectile then punctures the diaphragm
and enters the ram accelerator for further acceleration. Combustion is
initiated by shockwaves, stabilizes behind the projectile, and moves
Fig. 85. Schematic of an ODW [678]. δ is the wedge angle, θ is the flow angle, ε
is the oblique detonation wave angle.
with the projectile, sustaining a travelling pressure wave to accelerate
the projectile to high velocity. The principle of the ram accelerator is
similar to an airbreathing ramjet/scramjet engine, in which the sta­
design. The thrust, fuel-specific impulse, pressure, and frictional forces
tionary tube is analogous to the cylindrical outer cowling, and the
acting on the vehicles were evaluated, and it was shown that the
projectile is similar to the centerbody. The ram accelerator can be scaled
scramjet outperformed the shcramjet with a fuel-specific impulse of
up and down over a broad range of sizes because its operation principle
1450 s as opposed to 1109 s developed by the shcramjet. The authors
is independent of dimension. The ram accelerator can be readily used in
argued that the shcramjet had a combustor length of 1/5 of the scramjet
hypersonic test facilities for hypervelocity aerodynamics and re-entry
combustor length, was smaller and lighter than the scramjet, and
thermal protection and can be potentially of use in space launch of
required much less cooling than the scramjet. It is widely assumed that
insensitive components.
the shcramjet has a better overall propulsive performance than the
Stabilization of an ODW near the C-J condition has been the main
scramjet at higher Mach numbers, especially above Mach 12.
challenge to efficiently utilize such waves for propulsion in hypersonic
For shcramjet operation in the low-pressure, high altitude, high
regimes. Fig. 85 shows the typical structure of an oblique detonation
Mach number region, the reaction kinetics is of critical importance when
wave stabilized over a wedge [678]. An ODW is formed when the shock
maximizing scramjet performance because a high degree of chemical
and deflagration waves merge into a single structure. A slip line sepa­
dissociation will delay the completion of the exothermic reactions in the
rates the region of flow passing through the ODW and the deflagration
expansion nozzle, which is a large-loss mechanism deteriorating the
waves. The ODW driven by combustion heat release is steeper than the
production of thrust. Optimization of flowpath and exit nozzle shape for
wedge-induced oblique shock. When the normal component of the Mach
optimal shcramjet performance is required. Billig [117] recommended
number of the ODW is unity at a critical wedge angle δCJ, the detonation
that the engine should be operated at a fuel-rich equivalence ratio that
wave reaches the C-J point, at which the detonation wave is
was larger than required to provide adequate cooling at high hypersonic
self-sustained and corresponds to minimum entropy production. An

66
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

ODW engine or shcramjet is usually designed to operate close to C-J below 50 kPa, and a steady oblique detonation wave was generated only
conditions [119] over a wedge from flow conditions typical of a when the pressure condition was above the criteria for initiation. The
shcramjet combustor inlet, i.e., Mach ~7, a static pressure of ~23 kPa, authors divided the steady detonation wave into four kinds, i.e., strong
and a static temperature of ~750 K [678]. overdriven detonation wave, weak overdriven detonation wave,
To achieve satisfactory operation of shcramjets and ram accelerators, quasi-C-J detonation wave, and C-J detonation wave. Verreault and
practical issues that must be resolved are the stability of detonation Higgins [693,711] proposed energetic and kinetic limits to predict the
waves and conditions of C-J point over a range of combustor inlet ve­ conditions required to initiate an oblique detonation wave in a hydro­
locity, pressure, temperature, and equivalence ratio conditions [119]. gen/oxygen mixture. Experimental tests were performed to verify the
Extensive numerical studies have considered the impacting factors limits by varying the cone half angle and initial pressure in a stoichio­
leading to an unstable detonation, which include the wedge angle at metric hydrogen/oxygen mixture. Fig. 86 shows the five combustion
overdrive conditions [679,680], inhomogeneity of the incoming fuel-air regimes initiated by a conical projectile at various pressures and cone
mixture [681], effects of boundary layers [682,683], freestream turbu­ angles. Detailed explanations of energetic and kinetic limits were pro­
lence [684], and inflow unsteadiness [685]. The accuracy of those nu­ vided in Ref. [693]. Nevertheless, the results did not include effects of
merical studies on conditions achieving stable ODW depend on the projectile velocities on the ODW stability because the projectile velocity
complexity of the chemical models [686,687]. The stabilization over a was fixed at 1.5 times the C-J velocity.
wedge and the structure of ODW were experimentally observed using As such, shock-induced combustion experiments for hypersonic
PLIF in expansion tubes [688–690] and shock tube facilities [691,692]. propulsion are mostly focused on the fundamental analysis of the obli­
However, the unsteady behavior of the detonation structure was not que detonation wave, and the hypersonic vehicle driven by shcramjet
captured due to the short testing time of those hypersonic facilities. engines remains in the theoretical cycle analysis and numerical esti­
Aeroballistic testing, in which a projectile is fired into a stationary mation phase only. Further efforts on integrated experimental testing
combustible gas, was used to investigate shock-induced combustion and are needed though it is severely limited by the capability of replicating
oblique detonation due to its advantages over wind tunnel testing. the realistic flow conditions in the high Mach number regime.
Aeroballistic experiment is capable of testing in a full range of Mach
number and Reynolds numbers and free of contamination and boundary 4. Conclusion
layers effect occurring in hypersonic tunnels (shock tunnels, expansion
tubes, etc.) [693]. Numerous investigations of aeroballistic testing in a This paper contains a review of progress in supersonic combustion
combustible mixture used blunt projectiles to initiate shock-induced enhancement and stabilization for high Mach number propulsion.
combustion and detonation [694–701]. These experiments covering a Combustion stabilization is a critical problem for scramjets due to the
wide range of shock-induced combustion modes were used as test cases short flow residence time in the combustor, which is typically of the
for numerical schemes [702–708] to calculate hypersonic chemically order of a few milliseconds. During this time, fuel must be injected,
reacting flows [702]. Ess et al. [709] proposed use of blunt bodies, i.e., mixed with the incoming air, and burned efficiently before it is
cylindrical rods placed normal to the local flow direction in the expanded through the exhaust nozzle. Over the years, a wide range of
combustible flow for detonation wave generation in a propulsive duct of techniques aimed to maximize both mixing and combustion efficiencies
a shcramjet. Though aeroballistic tests of shock-induced combustion in a in a supersonic combustor have been proposed and tested. The large
wedge or a conical flow configuration are closer to the flowfields that variety of approaches used is justified by the fact that, in a typical
would occur in hypersonic ODW engines, experimental studies on mission, a scramjet engine maybe required to operate over a broad range
shock-induced combustion over wedge or conical hypersonic projectiles of conditions from take-over regime at lower altitude and lower speed
are not abundant. Kasahara et al. [710] explored the steady-state (Mach 4–5) to cruise conditions at higher altitude and speed in a range of
detonation structure generated around a hypersonic conical projectile Mach 7–10 or above. Over this regime, both stagnation enthalpy and
with a half cone angle of 120◦ in stoichiometric hydrogen-oxygen pre­ pressure change by some orders of magnitude, and therefore, the flow
mixed gases. The detonation waves were observed at initial pressures characteristics at the bottom and at the top of the range can vary and
need to be addressed with different techniques.
The first part of the paper was dedicated to a historical overview of
scramjet propulsion, its application and a summary of the overarching
problems and work performed. We have mainly divided the content of
the papers into diffusion combustion for low hypersonic regimes and
then partially or fully premixed combustion for high hypersonic re­
gimes. Each part is then further subdivided based on the flame stabili­
zation technique.
The first class of techniques focuses on the enhancement of mixing.
Combustion in scramjets is often defined as ‘mixing-controlled’ because
the time required to obtain good mixing is usually much longer than the
time required for ignition and complete combustion. Several approaches
to mixing enhancement have been investigated. A fundamental source of
mixing is provided by the fuel injectors. It has been demonstrated that an
efficient design of the injection system can significantly affect the mix­
ing properties of the flow. Shape of the orifices, jet momentum and di­
rection, and interaction among multiple jets have all been exploited to
promote turbulence and mixing in a scramjet combustor. Additional
hyper-mixers and vortex generators have been used in combination with
jets to enhance mixing. Oblique shocks have also been used to accelerate
the mixing layers growth rate by baroclinic effect. In these techniques,
the main purpose is providing a suitable environment to anchor and
stabilize the flame in proximity of the fuel injectors.
Fig. 86. Experimental results of observed combustion regimes along with A different approach is the use of cavity flameholders. At flow speeds
theoretical limits [693]. larger than the flame propagation speed, hypermixers and vortex

67
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

generation devices are not sufficient anymore to stabilize the flame. The performing the same measurements in different types of facilities under
flame is pushed away from the injectors and cannot sustain ignition of the same nominal conditions will help to understand and remove sys­
new propellant, leading to extinction (blowoff). In this case it is neces­ tematic facility bias, providing insights on extrapolating laboratory re­
sary to reduce the flow speed to increase the residence time inside the sults to actual flights. New opportunities are always presented with the
combustor. Cavities and steps provide a low speed recirculation region advent of novel diagnostics technologies, such as the proliferation of
where mixing can be completed, and combustion can be stabilized under pulse-burst laser systems in the past decade, finally allowing the appli­
suitable conditions. This stable flame can then be used to pilot and an­ cation of planar and volumetric imaging of the combustion phenomena
chor the main combustion in the freestream. At even larger flow speeds at frequencies suitable for hypersonic propulsion. Many more similar
the stagnation pressure losses introduced by cavities and other obstacles advances are expected in the coming decades and will bring new
in the flowfield become excessively detrimental for engine performance. experimental capabilities and new understanding. Simultaneously, ad­
In these cases, the use of plasma devices has been proposed as a com­ vances in high performance computing architecture, numerical algo­
bustion stabilization method. Plasma has the advantage of better rithms, and artificial intelligence will push our capabilities ever closer to
controllability over geometrical features; it can be turned on only when achieve high fidelity simulation of a full vehicle over different flight
necessary and tuned in order to minimize the negative impact on the scenarios. Last but not least, a fundamental contribution will need to be
flow. As described in Section 2.4, plasma has also been proposed in provided by flight testing as a final verification of the validity of the
several forms (plasma torches, laser induced plasma, non-equilibrium, design techniques adopted. Through these efforts, we anticipate a future
etc.) to provide flame ignition at flight speeds at which the tempera­ when supersonic combustor design will become routine, and standard­
ture recovery is insufficient for auto-ignition of the fuel. ized design guidelines can be implemented to achieve maximum
At the highest speed of interest for scramjet propulsion (Mach 8–10 combustor efficiency over the entire flight envelope.
or above), the residence time becomes even shorter, but the pressure
losses associated with cavities, struts or other devices becomes unac­
ceptable. At these Mach numbers, fuel injectors have to be inclined to a Declaration of competing interest
very shallow angle to reduce impact on the flow. At large Mach
numbers, the theoretical maximum specific impulse of a scramjet be­ The authors declare that they have no known competing financial
comes smaller and any possible source of thrust loss must be minimized. interests or personal relationships that could have appeared to influence
Since in this regime unstart is not a main concern, fuel is usually injected the work reported in this paper.
at the very beginning of the flowpath (inside the inlet) to allow more
time for mixing. At this location, the static temperature of the flow is still Acknowledgments
too low to produce significant combustion. Once the air-fuel mixture has
reached a sufficient degree of mixing, combustion is ignited by This work was generated in the process of working on U.S. Air Force
increasing the static temperature of the flow using an oblique shock, as Office of Scientific Research (AFOSR) grant FA9550-14-1-0343 grant
in the oblique detonation wave engines described in Section 3.2, or monitored by Dr. Chiping Li as part of the Presidential Early Career
using combined effects of alternating shocks and expansion regions, as Award for Scientists & Engineerings (PECASE) awarded to Dr. Tonghun
in the radical farming scramjets of Section 3.1. Lee. The authors gratefully acknowledge this source of support.
Despite the large amount of work performed, as summarized in this
review, supersonic combustion stabilization for hypersonic propulsion References
still presents formidable challenges for engineers. Although the overall
understanding of the physical process in a scramjet has been greatly [1] W. Heiser, D. Pratt, D. Daley, U. Mehta, Hypersonic Airbreathing Propulsion,
American Institute of Aeronautics and Astronautics, Inc., 1994, https://doi.org/
improved, the confidence in the currently available methods for quan­
10.2514/4.470356.
titative predictions is still very limited. Ground testing facilities are [2] D. Andreadis, Scramjets integrate air and space, Ind. Phys. 10 (2004) 26–29.
biased by technical difficulties. The flow produced by current hyper­ [3] H.-S. Tsien, Similarity laws of hypersonic flows, J. Math. Phys. 25 (1946)
247–251, https://doi.org/10.1002/sapm1946251247.
sonic wind tunnels is partially contaminated by impurities or solid
[4] J.D. Anderson, Hypersonic and High-Temperature Gas Dynamics, third ed.,
micro/nano particles that could affect the chemistry inside the American Institute of Aeronautics and Astronautics, Inc., Washington, DC, 2019
combustor. In addition, the air flow past the isolator is likely in a state of https://doi.org/10.2514/4.105142.
both vibrational and chemical non-equilibrium, that is different from the [5] R.L. Chase, M.H. Tang, A history of the NASP program from the formation of the
joint program Office to the termination of the HySTP scramjet performance
atmosphere and very difficult to characterize. Freestream turbulence is demonstration program, in: AIAA 6th Int. Aerosp. Planes Hypersonics Technol.
another variable that has been scarcely considered in the experimental Conf, 1995, https://doi.org/10.2514/6.1995-6031. AIAA 95-6031, Chattanooga,
work up to date, but that can have a dramatic impact on the results. TN.
[6] W.C. Engelund, Hyper-X aerodynamics: the X-43A airframe-integrated scramjet
Although much improved over the past few decades, diagnostic capa­ propulsion flight-test experiments, J. Spacecr. Rockets. 38 (2008) 801–802,
bilities for ground testing are still limited and affected by large uncer­ https://doi.org/10.2514/2.3757.
tainty. Laser and optical measurements have become routine in ground [7] W.C. Engelund, S.D. Holland, C.E. Cockrell, R.D. Bittner, Aerodynamic database
development for the hyper-X airframe-integrated scramjet propulsion
tests but still lack the spatial and temporal resolutions needed in hy­ experiments, J. Spacecr. Rockets. 38 (2008) 803–810, https://doi.org/10.2514/
personic flows. It is noticed that significant progress has been made in 2.3768.
the past few decades. Numerical simulations can potentially provide a [8] C. McClinton, A. Roudakov, V. Semenov, V. Kopehenov, Comparative flow path
analysis and design assessment of an axisymmetric hydrogen fueled scramjet
more comprehensive characterization of the flowfield, but great uncer­
flight test engine at a Mach number of 6.5, in: 7th Int. Sp. Planes Hypersonics
tainty still exists on the modeling of many fundamental phenomena such Syst. Technol. Conf, 1996, https://doi.org/10.2514/6.1996-4571. AIAA 96-4571,
as turbulent combustion, transport properties, chemical kinetics, Norfolk, VA.
[9] A.S. Roudakov, V.L. Semenov, V.I. Kopchenov, J.W. Hicks, B. Drive, Future flight
boundary layer transition, etc. Direct, ab-initio numerical simulations,
test plans of an axisymmetric hydrogen-fueled scramjet engine on the hypersonic
which can avoid or at least drastically limit physical models, represent flying laboratory, in: 7th Int. Spaceplanes Hypersonics Syst. Technol. Conf, 1996,
an appealing research tool, but the related computational costs will https://doi.org/10.2514/6.1996-4572. AIAA 96-4572, Norfolk, VA.
prevent these simulations from being practical for several more decades. [10] A. Roudakov, V. Semenov, J. Hicks, Recent flight test results of the joint CIAM-
NASA Mach 6.5 scramjet flight program, in: 8th AIAA Int. Sp. Planes Hypersonic
Future development in this field of research will require a closer Syst. Technol. Conf, 1998, https://doi.org/10.2514/6.1998-1643. AIAA 1998-
collaboration between experimental and computational tools. Mea­ 1643, Norfolk, VA.
surement and computational results can be used to validate each other [11] R. Voland, A. Auslender, M. Smart, A. Roudakov, V. Semenov, V. Kopchenov,
CIAM/NASA Mach 6.5 scramjet flight and ground test, in: 9th Int. Sp. Planes
and drastically reduce the overall uncertainty, increasing the level of Hypersonic Syst. Technol. Conf, 1999, https://doi.org/10.2514/6.1999-4848.
confidence of the engineers in the final design of a scramjet engine. Also AIAA 99-9848, Norfolk, VA.

68
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

[12] P. Liever, S. Habchi, W. Engelund, J. Martin, Stage separation analysis of the X- [37] S. Im, H. Do, Unstart phenomena induced by flow choking in scramjet inlet-
43A research vehicle, in: 22nd Appl. Aerodyn. Conf. Exhib, 2004, https://doi. isolators, Prog. Aerosp. Sci. 97 (2018) 1–21, https://doi.org/10.1016/j.
org/10.2514/6.2004-4725. AIAA 2004-4725, Providence, Rhode Island. paerosci.2017.12.001.
[13] C. Bahm, E. Baumann, J. Martin, D. Bose, R. Beck, B. Strovers, The X-43A hyper-X [38] J. Steelant, Achievements obtained for sustained hypersonic flight within the
mach 7 flight 2 guidance, navigation, and control overview and flight test results, LAPCAT project, in: 15th AIAA Int. Sp. Planes Hypersonic Syst. Technol. Conf,
in: AIAA/CIRA 13th Int. Sp. Planes Hypersonics Syst. Technol. Conf, 2005, 2008, https://doi.org/10.2514/6.2008-2578. AIAA 2008-2578, Dayton, Ohio.
https://doi.org/10.2514/6.2005-3275. AIAA 2005-3275, Capua, Italy. [39] J. Steelant, R. Varvill, C. Walton, S. Defoort, K. Hannemann, M. Marini,
[14] P. Harsha, L. Keel, A. Castrogiovanni, R. Sherrill, X-43A vehicle design and Achievements obtained for sustained hypersonic flight within the LAPCAT-II
manufacture, in: AIAA/CIRA 13th Int. Sp. Planes Hypersonics Syst. Technol. Conf, project, in: 20th AIAA Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2015,
2005, https://doi.org/10.2514/6.2005-3334. AIAA 2005-3334, Capua, Italy. https://doi.org/10.2514/6.2015-3677. AIAA 2015-3677, Glasgow, Scotland.
[15] E. Morelli, S. Derry, M. Smith, Aerodynamic parameter estimation for the X-43A [40] C. Meerts, J. Steelant, P. Hendrick, Preliminary design of the low speed
(Hyper-X) from flight data, in: AIAA Atmos. Flight Mech. Conf. Exhib, 2005, propulsion air intake of the lapcat-mr2 aircraft (ESA SP-692), ESA SP-692,
https://doi.org/10.2514/6.2005-5921. AIAA 2005-5921, San Francisco, Brugge, Belgium, in: Proc. 7th Eur. Symp. Aerothermodyn. Sp. Veh, 2011, p. 38,
California. http://articles.adsabs.harvard.edu/pdf/2011ESASP.692E.38M.
[16] C.W. Ohlhorst, D.E. Glass, W.E. Bruce, M.C. Lindell, W.L. Vaughn, R.W. Smith, R. [41] C. Meerts, J. Steelant, Air intake design for the dcceleration propulsion unit of the
B. Dirling, P.A. Hogenson, J.M. Nichols, N.W. Risner, D.R. Thompson, W. Kowbel, LAPCAT-MR2 hypersonic aircraft, in: 5th Eur. Conf. Aerosp. Sci, 2013. Munich,
B.J. Sullivan, J.R. Koenig, J.C. Cuneo, Development of X-43A mach 10 leading Germany.
edges, in: 56th Int. Astronaut. Congr. Int. Astronaut. Fed. Int. Acad. Astronaut. [42] J. Steelant, T. Langener, K. Hannemann, M. Marini, L. Serre, M. Bouchez,
Int. Inst. Sp. Law, 2005, https://doi.org/10.2514/6.IAC-05-D2.5.06. IAC-05- F. Falempin, Conceptual design of the high-speed propelled experimental flight
D2.5.06, Fukuoka, Japan. test vehicle HEXAFLY, in: 20th AIAA Int. Sp. Planes Hypersonic Syst. Technol.
[17] R.T. Voland, L.D. Huebner, C.R. McClinton, X-43A Hypersonic vehicle technology Conf, 2015, https://doi.org/10.2514/6.2015-3539. AIAA 2015-3539, Glasgow,
development, Acta Astronaut 59 (2006) 181–191, https://doi.org/10.1016/j. Scotland.
actaastro.2006.02.021. [43] J. Steelant, A. Mack, K. Hannemann, A. Gardner, Comparison of supersonic
[18] E. Baumann, C. Bahm, B. Strovers, R. Beck, M. Richard, The X-43A six degree of combustion tests with shock tunnels, flight and CFD, in: 42nd AIAA/ASME/SAE/
freedom Monte Carlo analysis, in: 46th AIAA Aerosp. Sci. Meet. Exhib, 2008, ASEE Jt. Propuls. Conf. Exhib, 2006, https://doi.org/10.2514/6.2006-4684.
https://doi.org/10.2514/6.2008-203. AIAA 2008-203, Reno, Nevada. AIAA 2006-4684, Sacramento, California.
[19] M.C. Davis, J.T. White, X-43A flight-test-determined aerodynamic force and [44] K. Hannemann, O. Haidn, G. Lamanna, I. Stotz, B. Weigand, J. Steelant, H. Ciezki,
moment characteristics at mach 7.0, J. Spacecr. Rockets 45 (2008) 472–484, M. Oschwald, S. Karl, J. Martinez Schramm, A. Mack, J. Sender, W. Clauss,
https://doi.org/10.2514/1.30413. C. Manfletti, Combustion experiments performed within the LAPCAT I project -
[20] Y. Lin, E. Baumann, D.M. Bose, R. Beck, G.D. Jenney, Tests and Techniques for an overview, in: 16th AIAA/DLR/DGLR Int. Sp. Planes Hypersonic Syst. Technol.
Characterizing and Modeling X-43A Electromechanical Actuators, 2008. http Conf, 2009, https://doi.org/10.2514/6.2009-7206. AIAA 2009-7206, Bremen,
s://ntrs.nasa.gov/citations/20090005180. Germany.
[21] S. Berry, K. Daryabeigi, K. Wurster, R. Bittner, Boundary-layer transition on X- [45] A. Vincent-Randonnier, Y. Moule, M. Ferrier, Combustion of hydrogen in hot air
43A, J. Spacecr. Rockets 47 (2010) 922–934, https://doi.org/10.2514/1.45889. flows within LAPCAT-II dual mode ramjet combustor at onera-LAERTE facility -
[22] C. Gibson, J. Neidhoefer, S. Cooper, L. Carlton, C. Cox, C. Jorgensen, experimental and numerical investigation, in: 19th AIAA Int. Sp. Planes
Development and flight test of the X-43A-LS hypersonic configuration UAV, in: Hypersonic Syst. Technol. Conf, 2014, https://doi.org/10.2514/6.2014-2932.
AIAA 1st Tech. Conf. Work. Unmanned Aerosp. Veh, 2002, https://doi.org/ AIAA 2014-2932, Atlanta, GA.
10.2514/6.2002-3462. AIAA 2002-3462, Portsmouth, Virginia. [46] P. Gruhn, A. Gülhan, Aerodynamic measurements of an air-breathing hypersonic
[23] D. Reubush, L. Nguyen, V. Rausch, Review of X-43A return to flight activities and vehicle at mach 3.5 to 8, AIAA J 56 (2018) 4282–4296, https://doi.org/10.2514/
current status, in: 12th AIAA Int. Sp. Planes Hypersonic Syst. Technol, 2003, 1.J056522.
https://doi.org/10.2514/6.2003-7085. AIAA 2003-7085, Norfolk, Virginia. [47] A. Vincent-Randonnier, V. Sabelnikov, A. Ristori, N. Zettervall, C. Fureby, An
[24] P. Joyce, J. Pomroy, L. Grindle, The hyper-X launch vehicle: challenges and experimental and computational study of hydrogen–air combustion in the
design considerations for hypersonic flight testing, in: AIAA/CIRA 13th Int. Sp. LAPCAT II supersonic combustor, Proc. Combust. Inst. 37 (2019) 3703–3711,
Planes Hypersonics Syst. Technol. Conf, 2005, https://doi.org/10.2514/6.2005- https://doi.org/10.1016/j.proci.2018.05.127.
3333. AIAA 2005-3333, Capua, Italy. [48] M.K. Smart, N.E. Hass, A. Paull, Flight data analysis of the HyShot 2 scramjet
[25] E. Baumann, Tailored excitation for frequency response measurement applied to flight experiment, AIAA J 44 (2006) 2366–2375, https://doi.org/10.2514/
the X-43A flight vehicle, in: 44th AIAA Aerosp. Sci. Meet. Exhib, 2006, https:// 1.20661.
doi.org/10.2514/6.2006-638. AIAA 2006-638, Reno, Nevada. [49] S. Walker, F. Rodgers, A. Paull, D. Van Wie, HyCAUSE flight test program, in:
[26] M. Davis, J. White, Flight-test-determined aerodynamic force and moment 15th AIAA Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2008, https://doi.org/
characteristics of the X-43A at mach 7.0, in: 14th AIAA/AHI Sp. Planes 10.2514/6.2008-2580. AIAA 2008-2580, Dayton, Ohio.
Hypersonic Syst. Technol. Conf, 2006, https://doi.org/10.2514/6.2006-8028. [50] S. Walker, F. Rodgers, The hypersonic collaborative Australia/United States
AIAA 2006-8028, Canberra, Australia. experiment (HyCAUSE), in: AIAA/CIRA 13th Int. Sp. Planes Hypersonics Syst.
[27] E. Baumann, J.W. Pahle, M.C. Davis, J.T. White, X-43A flush Airdata sensing Technol. Conf, 2005, https://doi.org/10.2514/6.2005-3254. AIAA 2005-3254,
system flight-test results, J. Spacecr. Rockets 47 (2010) 48–61, https://doi.org/ Capua, Italy.
10.2514/1.41163. [51] J. Larsson, S. Laurence, I. Bermejo-Moreno, J. Bodart, S. Karl, R. Vicquelin,
[28] J. Ellsworth, An analytical explanation for the X-43A flush air data sensing system Incipient thermal choking and stable shock-train formation in the heat-release
pressure mismatch between flight and theory, in: 28th AIAA Appl. Aerodyn. Conf, region of a scramjet combustor. Part II: large eddy simulations, Combust. Flame.
2010, https://doi.org/10.2514/6.2010-4964. AIAA 2010-4964, Chicago, Illinois. 162 (2015) 907–920, https://doi.org/10.1016/j.combustflame.2014.09.017.
[29] C.E. Cockrell, W.C. Engelund, R.D. Bittner, T.N. Jentink, A.D. Dilley, A. Frendi, [52] J. Larsson, Large eddy simulation of the HyShot II scramjet combustor using a
Integrated Aeropropulsive computational fluid dynamics methodology for the supersonic flamelet model, in: 48th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. &
hyper-X flight experiment, J. Spacecr. Rockets. 38 (2001) 836–843, https://doi. Exhib, 2012, https://doi.org/10.2514/6.2012-4261. AIAA 2012-4261, Reston,
org/10.2514/2.3773. Virigina.
[30] D. Freeman, D. Reubush, C. Mcclinton, V. Rausch, L. Crawford, The NASA hyper- [53] S.J. Laurence, D. Lieber, J. Martinez Schramm, K. Hannemann, J. Larsson,
X program, IAF-97-V.4.07, Turin, Italy, in: 48th Int. Astronaut. Congr, 1997, http Incipient thermal choking and stable shock-train formation in the heat-release
s://ntrs.nasa.gov/citations/19980018480. region of a scramjet combustor. Part I: shock-tunnel experiments, Combust.
[31] P.L. Moses, V.L. Rausch, L.T. Nguyen, J.R. Hill, NASA hypersonic flight Flame. 162 (2015) 921–931, https://doi.org/10.1016/j.
demonstrators—overview, status, and future plans, Acta Astronaut 55 (2004) combustflame.2014.09.016.
619–630, https://doi.org/10.1016/j.actaastro.2004.05.045. [54] R. Pecnik, V.E. Terrapon, F. Ham, G. Iaccarino, H. Pitsch, Reynolds-averaged
[32] S. Ferlemann, C. McClinton, K. Rock, R. Voland, Hyper-X mach 7 scramjet design, Navier-Stokes simulations of the HyShot II scramjet, AIAA J 50 (2012)
ground test and flight results, in: AIAA/CIRA 13th Int. Sp. Planes Hypersonics 1717–1732, https://doi.org/10.2514/1.J051473.
Syst. Technol. Conf, 2005, https://doi.org/10.2514/6.2005-3322. AIAA 2005- [55] M. Chapuis, E. Fedina, C. Fureby, K. Hannemann, S. Karl, J. Martinez Schramm,
3322, Capua, Italy. A computational study of the HyShot II combustor performance, Proc. Combust.
[33] L. Marshall, C. Bahm, G. Corpening, R. Sherrill, Overview with results and lessons Inst. 34 (2013) 2101–2109, https://doi.org/10.1016/j.proci.2012.07.014.
learned of the X-43A mach 10 flight, in: AIAA/CIRA 13th Int. Sp. Planes [56] K. Nordin-Bates, C. Fureby, S. Karl, K. Hannemann, Understanding scramjet
Hypersonics Syst. Technol. Conf, 2005, https://doi.org/10.2514/6.2005-3336. combustion using LES of the HyShot II combustor, Proc. Combust. Inst. 36 (2017)
AIAA 2005-3336, Capua, Italy. 2893–2900, https://doi.org/10.1016/j.proci.2016.07.118.
[34] C. McClinton, X-43-Scramjet power breaks the hypersonic barrier: dryden [57] K. Nordin-Bates, C. Fureby, Understanding scramjet combustion using LES of the
lectureship in research for 2006, in: 44th AIAA Aerosp. Sci. Meet. Exhib, 2006, HyShot II combustor: stable combustion and incipient thermal choking, in: 51st
https://doi.org/10.2514/6.2006-1. AIAA 2006-1, Reno, Nevada. AIAA/SAE/ASEE Jt. Propuls. Conf, 2015, https://doi.org/10.2514/6.2015-3838.
[35] J. Hank, Air force research laboratory hypersonic propulsion research programs, AIAA 2015-3838, Orlando, FL.
in: 43rd AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2007, https://doi.org/ [58] M.R. Gruber, K. Jackson, J. Liu, Hydrocarbon-fueled Scramjet Combustor
10.2514/6.2007-5371. AIAA 2007-5371, Cincinnati, OH. Flowpath Development for Mach 6-8 HIFiRE Flight Experiments, Wright-
[36] J. Hank, J. Murphy, R. Mutzman, The X-51A scramjet engine flight demonstration Patterson Air Force Base, OH, 2008. https://apps.dtic.mil/sti/pdfs/ADA532732.
program, in: 15th AIAA Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2008, pdf.
https://doi.org/10.2514/6.2008-2540. AIAA 2008-2540, Dayton, Ohio. [59] A. Auslender, K. Suder, S. Thomas, An overview of the NASA FAP hypersonics
project airbreathing propulsion research, in: 16th AIAA/DLR/DGLR Int. Sp.

69
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Planes Hypersonic Syst. Technol. Conf, 2009, https://doi.org/10.2514/6.2009- [85] B. McGann, T. Lee, T. Ombrello, C.D. Carter, S.D. Hammack, H. Do, Inlet
7277. AIAA 2009-7277, Bremen, Germany. distortion effects on fuel distribution and ignition in scramjet cavity flameholder,
[60] K. Jackson, M. Gruber, T. Barhorst, The HIFiRE flight 2 experiment: an overview J. Propuls. Power. 35 (2019) 601–613, https://doi.org/10.2514/1.B37204.
and status update, in: 45th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, [86] M.R. Gruber, A.S. Nejad, T.H. Chen, J.C. Dutton, Compressibility effects in
2009, https://doi.org/10.2514/6.2009-5029. AIAA 2009-5029, Denver, supersonic transverse injection flowfields, Phys. Fluids. 9 (1997) 1448–1461,
Colorado. https://doi.org/10.1063/1.869257.
[61] K.R. Jackson, M.R. Gruber, S. Buccellato, Mach 6-8+ hydrocarbon-fueled [87] M.R. Gruber, R.A. Baurle, T. Mathur, K.-Y. Hsu, Fundamental studies of cavity-
scramjet flight experiment: the HIFiRE flight 2 project, J. Propuls. Power. 31 based flameholder concepts for supersonic combustors, J. Propuls. Power. 17
(2015) 36–53, https://doi.org/10.2514/1.B35350. (2001) 146–153, https://doi.org/10.2514/2.5720.
[62] K. Jackson, M. Gruber, S. Buccellato, HIFiRE flight 2 - a program overview, in: [88] S.B. Leonov, A. Houpt, B. Hedlund, Experimental demonstration of plasma-based
51st AIAA Aerosp. Sci. Meet. Incl. New Horizons Forum Aerosp. Expo, 2013, flameholder in a model scramjet, in: 21st AIAA Int. Sp. Planes Hypersonics
https://doi.org/10.2514/6.2013-695. AIAA 2013-0695, Grapevine (Dallas/Ft. Technol. Conf, 2017, https://doi.org/10.2514/6.2017-2249. AIAA 2017-2249,
Worth Region), Texas. Xiamen, China.
[63] K. Bowcutt, A. Paull, D. Dolvin, M. Smart, HIFiRE: an international collaboration [89] J.A. Fulton, J.R. Edwards, A. Cutler, J. McDaniel, C. Goyne, Turbulence/
to advance the science and technology of hypersonic flight. 28th Int. Congr. chemistry interactions in a ramp-stabilized supersonic hydrogen–air diffusion
Aeronaut. Sci., International Council of the Aeronautical Sciences (ICAS), flame, Combust. Flame. 174 (2016) 152–165, https://doi.org/10.1016/j.
Brisbane, QLD, Australia, 2012. combustflame.2016.09.017.
[64] K. Jackson, M. Gruber, S. Buccellato, HIFiRE flight 2 project overview and status [90] D. Baccarella, Q. Liu, A. Passaro, T. Lee, H. Do, Development and testing of the
update 2011, in: 17th AIAA Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2011, ACT-1 experimental facility for hypersonic combustion research, Meas. Sci.
https://doi.org/10.2514/6.2011-2202. AIAA 2011-2202, San Francisco, Technol. 27 (2016), 045902, https://doi.org/10.1088/0957-0233/27/4/045902.
California. [91] D. Baccarella, Q. Liu, T. Lee, S.D. Hammack, H. Do, The supersonic combustion
[65] M. Roberts, M. Smart, M. Frost, HIFiRE 7: design to achieve scientific goals, in: facility ACT-2, in: 55th AIAA Aerosp. Sci. Meet, 2017, https://doi.org/10.2514/
18th AIAA/3AF Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2012, https://doi. 6.2017-0103. AIAA 2017-0103, Grapevine, Texas.
org/10.2514/6.2012-5841. AIAA 2012-5841, Tours, France. [92] Z. Jiang, H. Yu, Experiments and development of long-test-duration hypervelocity
[66] M. Bolender, J. Staines, D. Dolvin, HIFiRE 6: an adaptive flight control detonation-driven shock tunnel (LHDst), in: 52nd Aerosp. Sci. Meet, 2014,
experiment, in: 50th AIAA Aerosp. Sci. Meet. Incl. New Horizons Forum Aerosp. https://doi.org/10.2514/6.2014-1012. AIAA 2014-1012, National Harbor,
Expo, 2012, https://doi.org/10.2514/6.2012-252. AIAA 2012-0252, Nashville, Maryland.
Tennessee. [93] Z. Jiang, H. Yu, Successful development of the long-test-duration hypervelocity
[67] D.E. Glass, D. Capriotti, T. Reimer, M. Kütemeyer, M. Smart, Testing of DLR C/C- detonation-driven shock tunnel, in: R. Bonazza, D. Ranjan (Eds.), 29th Int. Symp.
SiC and C/C for HIFiRE 8 scramjet combustor, in: 19th AIAA Int. Sp. Planes Shock Waves 1, Springer International Publishing, 2013, pp. 51–58, https://doi.
Hypersonic Syst. Technol. Conf, 2014, https://doi.org/10.2514/6.2014-3089. org/10.1007/978-3-319-16835-7.
AIAA 2014-3089, Atlanta, GA. [94] Corin Segal, The Scramjet Engine: Processes and Characteristics, Cambridge
[68] H. Alesi, A. Paull, M. Smart, K.G. Bowcutt, A concept for the HIFiRE 8 flight test, University Press, New York, NY, 2009.
Tromso, Norway, in: L. Ouwehand (Ed.), 22nd ESA Symp. Eur. Rocket Balloon [95] H. Do, S. Im, M.G. Mungal, M.A. Cappelli, Visualizing supersonic inlet duct
Program. Relat. Res, 2015, http://articles.adsabs.harvard.edu/pdf/2015ESASP unstart using planar laser Rayleigh scattering, Exp. Fluids. 50 (2011) 1651–1657,
.730.401A. https://doi.org/10.1007/s00348-010-1028-4.
[69] A.F. Woolf, Conventional Prompt Global Strike and Long-Range Ballistic Missiles: [96] Z. Wang, H. Wang, M. Sun, Review of cavity-stabilized combustion for scramjet
Background and Issues, DIANE Publishing, 2012. applications, Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng. 228 (2014) 2718–2735,
[70] S.L. Gai, Free piston shock tunnels: developments and capabilities, Prog. Aerosp. https://doi.org/10.1177/0954410014521172.
Sci. 29 (1992) 1–41, https://doi.org/10.1016/0376-0421(92)90002-Y. [97] F.W. Barnes, C. Segal, Cavity-based flameholding for chemically-reacting
[71] F.K. Lu, D.E. Marren, Advanced Hypersonic Test Facilities, vol. 198, American supersonic flows, Prog. Aerosp. Sci. 76 (2015) 24–41, https://doi.org/10.1016/j.
Institute of Aeronautics and Astronautics, Reston ,VA, 2002, https://doi.org/ paerosci.2015.04.002.
10.2514/4.866678. [98] M.R. Gruber, A.S. Nejadt, T.H. Chen, J.C. Dutton, Mixing and penetration studies
[72] S. Schneider, Design of a Mach-6 quiet-flow wind-tunnel nozzle using the e**N of sonic jets in a mach 2 freestream, J. Propuls. Power. 11 (1995) 315–323,
method for transition estimation, in: 36th AIAA Aerosp. Sci. Meet. Exhib, 1998, https://doi.org/10.2514/3.51427.
https://doi.org/10.2514/6.1998-547. AIAA 1998-547, Reno, NV. [99] A. Ben-Yakar, R. Hanson, Ultra-fast-framing schlieren system for studies of the
[73] S.P. Schneider, Development of hypersonic quiet tunnels, J. Spacecr. Rockets. 45 time evolution of jets in supersonic crossflows, Exp. Fluids. 32 (2002) 652–666,
(2008) 641–664, https://doi.org/10.2514/1.34489. https://doi.org/10.1007/s00348-002-0405-z.
[74] J. Hofferth, R. Bowersox, W. Saric, The mach 6 quiet tunnel at Texas A&M: quiet [100] A. Ben-Yakar, M.G. Mungal, R.K. Hanson, Time evolution and mixing
flow performance, in: 27th AIAA Aerodyn. Meas. Technol. Gr. Test. Conf, 2010, characteristics of hydrogen and ethylene transverse jets in supersonic crossflows,
https://doi.org/10.2514/6.2010-4794. AIAA 2010-4794, Chicago, Illinois. Phys. Fluids. 18 (2006), 026101, https://doi.org/10.1063/1.2139684.
[75] M.T. Lakebrink, K.G. Bowcutt, T. Winfree, C.C. Huffman, T.J. Juliano, [101] A.S. Potturi, J.R. Edwards, Large-eddy/Reynolds-averaged Navier–Stokes
Optimization of a mach-6 quiet wind-tunnel nozzle, J. Spacecr. Rockets. 55 simulation of cavity-stabilized ethylene combustion, Combust. Flame. 162 (2015)
(2018) 315–321, https://doi.org/10.2514/1.A33794. 1176–1192, https://doi.org/10.1016/j.combustflame.2014.10.011.
[76] A. Ben-Yakar, R.K. Hanson, Characterization of expansion tube flows for [102] B.E. Milton, K. Pianthong, Pulsed, supersonic fuel jets—a review of their
hypervelocity combustion studies, J. Propuls. Power. 18 (2002) 943–952, https:// characteristics and potential for fuel injection, Int. J. Heat Fluid Flow. 26 (2005)
doi.org/10.2514/2.6021. 656–671, https://doi.org/10.1016/j.ijheatfluidflow.2005.03.002.
[77] A. Dufrene, M. Sharma, J.M. Austin, Design and characterization of a [103] T. Ombrello, C. Carter, J. McCall, F. Schauer, A. Naples, J. Hoke, K.-Y. Hsu,
hypervelocity expansion tube facility, J. Propuls. Power. 23 (2007) 1185–1193, Enhanced mixing in supersonic flow using a pulse detonator, J. Propuls. Power.
https://doi.org/10.2514/1.30349. 31 (2015) 654–663, https://doi.org/10.2514/1.B35316.
[78] Y.M. Abul-Huda, M. Gamba, Flow characterization of a hypersonic expansion [104] S. Leonov, D. Yarantsev, C. Carter, Experiments on electrically controlled
tube facility for supersonic combustion studies, J. Propuls. Power. 33 (2017) flameholding on a plane wall in a supersonic airflow, J. Propuls. Power. 25 (2009)
1504–1519, https://doi.org/10.2514/1.B36543. 289–294, https://doi.org/10.2514/1.38002.
[79] K. Hannemann, High enthalpy flows in the HEG shock tunnel: experiment and [105] S. Leonov, D. Yarantsev, V. Sabelnikov, Electrically driven combustion near the
numerical rebuilding, in: 41st Aerosp. Sci. Meet. Exhib, 2003, https://doi.org/ plane wall in a supersonic duct, in: Prog. Propuls. Phys, 2011, pp. 519–530,
10.2514/6.2003-978. AIAA 2003-978, Reno, Nevada. https://doi.org/10.1051/eucass/201102519. EDP Sciences, Les Ulis, France.
[80] R.J. Stalker, A. Paull, D. Mee, R.G. Morgan, P.A. Jacobs, Scramjets and shock [106] J.F. Driscoll, C.C. Rasmussen, Correlation and analysis of blowout limits of flames
tunnels—the Queensland experience, Prog. Aerosp. Sci. 41 (2005) 471–513, in high-speed airflows, J. Propuls. Power. 21 (2005) 1035–1044, https://doi.org/
https://doi.org/10.1016/j.paerosci.2005.08.002. 10.2514/1.13329.
[81] D.J. Mee, R.G. Morgan, A. Paull, P.A. Jacobs, M.K. Smart, The T4 stalker tube, in: [107] A. Vincent-Randonnier, S.B. Leonov, D. Packan, First experiments on plasma
G. Ben-Dor, O. Sadot, O. Igra (Eds.), 30th Int. Symp. Shock Waves 2, Springer assisted supersonic combustion at LAERTE facility, in: 55th AIAA Aerosp. Sci.
International Publishing, 2017, pp. 1425–1429, https://doi.org/10.1007/978-3- Meet, 2017, https://doi.org/10.2514/6.2017-1975. AIAA 2017-1975, Grapevine,
319-46213-4. Texas.
[82] R.G. Morgan, D.E. Gildfind, X3 reflected shock tunnel for extended flow duration, [108] A. Houpt, S. Leonov, T. Ombrello, C. Carter, R.J. Leiweke, Flow control in
in: H. Haiyan (Ed.), 2014 Asia-Pacific Int. Symp. Aerosp. Technol, 2014, pp. 1–12. supersonic-cavity-based airflow by quasi-direct-current electric discharge, AIAA J
Shanghai, China. 57 (2019) 2881–2891, https://doi.org/10.2514/1.J058057.
[83] B. McGann, C.D. Carter, T. Ombrello, H. Do, Direct spectrum matching of laser- [109] G. Masuya, B. Choi, N. Ichikawa, K. Takita, Mixing and combustion of fuel jet in
induced breakdown for concentration and gas density measurements in turbulent pseudo-shock waves, in: 40th AIAA Aerosp. Sci. Meet. Exhib, 2002, https://doi.
reacting flows, Combust. Flame. 162 (2015) 4479–4485, https://doi.org/ org/10.2514/6.2002-809. AIAA 2002-0809, Reno, NV.
10.1016/j.combustflame.2015.08.021. [110] A.Y. Starikovskii, Plasma supported combustion, Proc. Combust. Inst. 30 (2005)
[84] H. Do, C.D. Carter, Q. Liu, T.M. Ombrello, S. Hammack, T. Lee, K.-Y. Hsu, 2405–2417, https://doi.org/10.1016/j.proci.2004.08.272.
Simultaneous gas density and fuel concentration measurements in a supersonic [111] T. Ombrello, X. Qin, Y. Ju, A. Gutsol, A. Fridman, C. Carter, Combustion
combustor using laser induced breakdown, Proc. Combust. Inst. 35 (2015) enhancement via stabilized piecewise nonequilibrium gliding arc plasma
2155–2162, https://doi.org/10.1016/J.PROCI.2014.07.043. discharge, AIAA J 44 (2006) 142–150, https://doi.org/10.2514/1.17018.

70
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

[112] H. Do, M.G. Mungal, M.A. Cappelli, Jet flame ignition in a supersonic crossflow [140] T. Mitani, T. Kouchi, Flame structures and combustion efficiency computed for a
using a pulsed nonequilibrium plasma discharge, IEEE Trans. Plasma Sci. 36 Mach 6 scramjet engine, Combust. Flame. 142 (2005) 187–196, https://doi.org/
(2008) 2918–2923, https://doi.org/10.1109/TPS.2008.2005986. 10.1016/j.combustflame.2004.10.004.
[113] S. Keshav, Y.G. Utkin, M. Nishihara, A. Bao, J.W. Rich, I.V. Adamovich, Studies of [141] J.A. Fulton, J.R. Edwards, H.A. Hassan, J.C. McDaniel, C.P. Goyne, R.D. Rockwell,
chemi-ionization and chemiluminescence in supersonic flows of combustion A.D. Cutler, C.T. Johansen, P.M. Danehy, Large-Eddy/Reynolds-Averaged
products, J. Thermophys. Heat Transf. 22 (2008) 157–167, https://doi.org/ Navier–Stokes simulations of reactive flow in dual-mode scramjet combustor, J.
10.2514/1.30822. Propuls. Power. 30 (2014) 558–575, https://doi.org/10.2514/1.B34929.
[114] S. Leonov, C. Savelkin, D. Yarantsev, C. Carter, V. Sermanov, M. Starodubtsev, [142] Y. Moule, V. Sabel’nikov, A. Mura, M. Smart, Computational fluid dynamics
Experiments on plasma-assisted combustion in M=2 hot test-bed PWT-50H, in: investigation of a mach 12 scramjet engine, J. Propuls. Power. 30 (2014)
46th AIAA Aerosp. Sci. Meet. Exhib, 2008, https://doi.org/10.2514/6.2008- 461–473, https://doi.org/10.2514/1.B34992.
1359. AIAA 2008-1359, Reno, Nevada. [143] W.O. Landsberg, V. Wheatley, M.K. Smart, A. Veeraragavan, Enhanced supersonic
[115] A. Firsov, K.V. Savelkin, D.A. Yarantsev, S.B. Leonov, Plasma-enhanced mixing combustion targeting combustor length reduction in a mach 12 scramjet, AIAA J
and flameholding in supersonic flow, Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 56 (2018) 3802–3807, https://doi.org/10.2514/1.J057417.
373 (2015) 20140337, https://doi.org/10.1098/rsta.2014.0337. [144] D. Curran, V. Wheatley, M. Smart, Investigation of combustion mode control in a
[116] Y. Ju, W. Sun, Plasma assisted combustion: dynamics and chemistry, Prog. Energy mach 8 shape-transitioning scramjet, AIAA J 57 (2019) 2977–2988, https://doi.
Combust. Sci. 48 (2015) 21–83, https://doi.org/10.1016/j.proci.2014.05.021. org/10.2514/1.J057999.
[117] F.S. Billig, Research on supersonic combustion, J. Propuls. Power. 9 (1993) [145] Z.J. Denman, S. Brieschenk, A. Veeraragavan, V. Wheatley, M. Smart,
499–514, https://doi.org/10.2514/3.23652. Experimental design of a cavity flameholder in a mach 8 shape-transitioning
[118] J. Odam, Scramjet Experiments Using Radical Farming, The University of scramjet, in: 19th AIAA Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2014,
Queensland, 2004. https://doi.org/10.2514/6.2014-2953. AIAA 2014-2953, Atlanta, GA.
[119] R. Dudebout, J.P. Sislian, R. Oppitz, Numerical simulation of hypersonic shock- [146] M. Gamba, M.G. Mungal, Ignition, flame structure and near-wall burning in
induced combustion ramjets, J. Propuls. Power. 14 (1998) 869–879, https://doi. transverse hydrogen jets in supersonic crossflow, J. Fluid Mech. 780 (2015)
org/10.2514/2.5368. 226–273, https://doi.org/10.1017/jfm.2015.454.
[120] R. Dunlap, A preliminary study of the application of steady-state detonative [147] A.R. Karagozian, Transverse jets and their control, Prog. Energy Combust. Sci. 36
combustion to a reaction engine, J. Jet Propuls. 28 (1958) 451–456, https://doi. (2010) 531–553, https://doi.org/10.1016/j.pecs.2010.01.001.
org/10.2514/8.7347. [148] K. Mahesh, The interaction of jets with crossflow, Annu. Rev. Fluid Mech. 45
[121] W.H. Sargent, Detonation wave hypersonic ramjet, ARS J 30 (1960) 543–549, (2013) 379–407, https://doi.org/10.1146/annurev-fluid-120710-101115.
https://doi.org/10.2514/8.5147. [149] M. Sun, S. Zhang, Y. Zhao, Y. Zhao, J. Liang, Experimental investigation on
[122] D.T. Pratt, J.W. Humphrey, D.E. Glenn, Morphology of standing oblique transverse jet penetration into a supersonic turbulent crossflow, Sci. China
detonation waves, J. Propuls. Power. 7 (1991) 837–845, https://doi.org/ Technol. Sci. 56 (2013) 1989–1998, https://doi.org/10.1007/s11431-013-5265-
10.2514/3.23399. 7.
[123] J.P. Sislian, Detonation-wave ramjets, in: S.N.B. Murthy, E.T. Curran (Eds.), [150] E. Zukoski, F.W. Spaid, Secondary injection of gases into a supersonic flow, AIAA
Scramjet Propuls, American Institute of Aeronautics and Astronautics, Reston , J 2 (1964) 1689–1696, https://doi.org/10.2514/3.2653.
VA, 2001, pp. 823–889, https://doi.org/10.2514/5.9781600866609.0823.0889. [151] L.J. Chrans, D.J. Colins, Stagnation temperature and molecular weight effects in
[124] X. Fan, G. Yu, J. Li, X. Zhang, C.-J. Sung, Investigation of vaporized kerosene jet interaction, AIAA J 8 (1970) 584–585, https://doi.org/10.2514/3.5718.
injection and combustion in a supersonic model combustor, J. Propuls. Power. 22 [152] M. Gamba, V.A. Miller, M.G. Mungal, R.K. Hanson, Temperature and number
(2006) 103–110, https://doi.org/10.2514/1.15427. density measurement in non-uniform supersonic flowfields undergoing mixing
[125] J. Beloki Perurena, C.O. Asma, R. Theunissen, O. Chazot, Experimental using toluene PLIF thermometry, Appl. Phys. B. 120 (2015) 285–304, https://doi.
investigation of liquid jet injection into Mach 6 hypersonic crossflow, Exp. Fluids. org/10.1007/s00340-015-6136-7.
46 (2009) 403–417, https://doi.org/10.1007/s00348-008-0566-5. [153] C.R. McClinton, Effect of Ratio of Wall Boundary-layer Thickness to Jet Diameter
[126] P. Li, Z. Wang, X.-S. Bai, H. Wang, M. Sun, L. Wu, C. Liu, Three-dimensional flow on Mixing of a Normal Hydrogen Jet in a Supersonic Stream, Hampton, VA, 1974.
structures and droplet-gas mixing process of a liquid jet in supersonic crossflow, https://ntrs.nasa.gov/citations/19740018713.
Aerosp. Sci. Technol. 90 (2019) 140–156, https://doi.org/10.1016/j. [154] R. Portz, C. Segal, Penetration of gaseous jets in supersonic flows, AIAA J 44
ast.2019.04.024. (2006) 2426–2429, https://doi.org/10.2514/1.23541.
[127] F. Xiao, M.B. Sun, Effects of Mach number on liquid jet primary breakup in gas [155] W. Heltsley, J. Snyder, C. Cheung, M. Mungal, R. Hanson, Combustion stability
crossflow, At. Sprays. 28 (2018) 975–999, https://doi.org/10.1615/ regimes of hydrogen jets in supersonic crossflow, in: 43rd AIAA/ASME/SAE/
ATOMIZSPR.2019026846. ASEE Jt. Propuls. Conf. Exhib, 2007, https://doi.org/10.2514/6.2007-5401.
[128] X. Fan, G. Yu, J. Li, X. Lu, X. Zhang, C.-J. Sung, Combustion and ignition of AIAA 2007-5401, Cincinnati, OH.
thermally cracked kerosene in supersonic model combustors, J. Propuls. Power. [156] J.A. Schetz, P.F. Hawkins, H. Lehman, Structure of highly underexpanded
23 (2007) 317–324, https://doi.org/10.2514/1.26402. transverse jets in a supersonic stream, AIAA J 5 (1967) 882–884, https://doi.org/
[129] Y. Tian, S. Yang, J. Le, F. Zhong, X. Tian, Investigation of combustion process of a 10.2514/3.4095.
kerosene fueled combustor with air throttling, Combust. Flame. 179 (2017) [157] J.A. Schetz, R.A. Weinraub, R.E. Mahaffey, Supersonic transverse injection into a
74–85, https://doi.org/10.1016/j.combustflame.2017.01.021. supersonic stream, AIAA J 6 (1968) 933–934, https://doi.org/10.2514/3.4631.
[130] J. Urzay, Supersonic combustion in air-breathing propulsion systems for [158] R.C. Orth, J.A. Funk, An experimental and comparative study of jet penetration in
hypersonic flight, Annu. Rev. Fluid Mech. 50 (2018) 593–627, https://doi.org/ supersonic flow, J. Spacecr. Rockets. 4 (1967) 1236–1242, https://doi.org/
10.1146/annurev-fluid-122316-045217. 10.2514/3.29058.
[131] J. Chang, J. Zhang, W. Bao, D. Yu, Research progress on strut-equipped [159] F.W. Spaid, E.E. Zukoski, A study of the interaction of gaseous jets from transverse
supersonic combustors for scramjet application, Prog. Aerosp. Sci. 103 (2018) slots with supersonic external flows, AIAA J 6 (1968) 205–212, https://doi.org/
1–30, https://doi.org/10.1016/j.paerosci.2018.10.002. 10.2514/3.4479.
[132] A. Ferri, Mixing-controlled supersonic combustion, Annu. Rev. Fluid Mech. 5 [160] R.C. Orth, J.A. Schetz, F.S. Billig, The Interaction and Penetration of Gaseous Jets
(1973) 301–338, https://doi.org/10.1146/annurev.fl.05.010173.001505. in Supersonic Flow, 1969. Washington, DC.
[133] E.T. Curran, W.H. Heiser, D.T. Pratt, Fluid phenomena in scramjet combustion [161] F.S. Billig, R.C. Orth, M. Lasky, A unified analysis of gaseous jet penetration, AIAA
systems, Annu. Rev. Fluid Mech. 28 (1996) 323–360, https://doi.org/10.1146/ J 9 (1971) 1048–1058, https://doi.org/10.2514/3.49916.
annurev.fl.28.010196.001543. [162] L.S. Cohen, L.J. Coulter, W.J. Egan, Penetration and mixing of multiple gas jets
[134] E.D. Gonzalez-Juez, A.R. Kerstein, R. Ranjan, S. Menon, Advances and challenges subjected to a cross flow, AIAA J 9 (1971) 718–724, https://doi.org/10.2514/
in modeling high-speed turbulent combustion in propulsion systems, Prog. Energy 3.6253.
Combust. Sci. 60 (2017) 26–67, https://doi.org/10.1016/j.pecs.2016.12.003. [163] R.C. Rogers, A Study of the Mixing of Hydrogen Injected Normal to a Supersonic
[135] R. Yentsch, D. Gaitonde, Numerical investigation of the HIFiRE-2 scramjet Airstream, 1971. Hampton, VA.
flowpath, in: 51st AIAA Aerosp. Sci. Meet. Incl. New Horizons Forum Aerosp. [164] R.C. Rogers, Mixing of Hydrogen Injected from Multiple Injectors Normal to a
Expo, 2013, https://doi.org/10.2514/6.2013-119. AIAA 2013-0119, Grapevine Supersonic Airstream, 1971. Hampton, VA.
(Dallas/Ft. Worth Region), Texas. [165] J.G. Santiago, J.C. Dutton, Velocity measurements of a jet injected into a
[136] R.J. Yentsch, D.V. Gaitonde, Numerical investigation of dual-mode operation in a supersonic crossflow, J. Propuls. Power. 13 (1997) 264–273, https://doi.org/
rectangular scramjet flowpath, J. Propuls. Power. 30 (2014) 474–489, https:// 10.2514/2.5158.
doi.org/10.2514/1.B34994. [166] E. Erdem, S. Saravanan, J. Lin, K. Kontis, Experimental investigation of transverse
[137] R.J. Yentsch, D.V. Gaitonde, Unsteady three-dimensional mode transition injection flowfield at march 5 and the influence of impinging shock wave, in: 18th
phenomena in a scramjet flowpath, J. Propuls. Power. 31 (2015) 104–122, AIAA/3AF Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2012, https://doi.org/
https://doi.org/10.2514/1.B35205. 10.2514/6.2012-5800. AIAA 2012-5800, Tours, France.
[138] A. Saghafian, L. Shunn, D.A. Philips, F. Ham, Large eddy simulations of the [167] Q. Liu, D. Baccarella, B. McGann, T. Lee, H. Do, Experimental investigation of
HIFiRE scramjet using a compressible flamelet/progress variable approach, Proc. single jet and dual jet injection in a supersonic combustor, in: 2018 AIAA Aerosp.
Combust. Inst. 35 (2015) 2163–2172, https://doi.org/10.1016/j. Sci. Meet, 2018, https://doi.org/10.2514/6.2018-1363. AIAA 2018-1363,
proci.2014.10.004. Kissimmee, Florida.
[139] D. Cecere, A. Ingenito, E. Giacomazzi, L. Romagnosi, C. Bruno, Hydrogen/air [168] K. Lin, M. Ryan, C. Carter, M. Gruber, C. Raffoul, Scalability of ethylene gaseous
supersonic combustion for future hypersonic vehicles, Int. J. Hydrogen Energy. 36 jets for fueling high-speed air-breathing combustors, in: 47th AIAA Aerosp. Sci.
(2011) 11969–11984, https://doi.org/10.1016/j.ijhydene.2011.06.051. Meet. Incl. New Horizons Forum Aerosp. Expo, 2009, https://doi.org/10.2514/
6.2009-1423. AIAA 2009-1423, Orlando, Florida.

71
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

[169] K. Lin, M. Ryan, C. Carter, M. Gruber, C. Raffoul, Raman scattering measurements [196] M. Zhao, T. Ye, C. Cao, T. Zhou, M. Zhu, Study of sonic injection from circular
of gaseous ethylene jets in a mach 2 supersonic crossflow, J. Propuls. Power. 26 injector into a supersonic cross-flow using large eddy simulation, Int. J. Hydrogen
(2010) 503–513, https://doi.org/10.2514/1.43757. Energy. 41 (2016) 17657–17669, https://doi.org/10.1016/j.
[170] J.C. McDaniel, J. Graves, Laser-induced-fluorescence visualization of transverse ijhydene.2016.07.018.
gaseous injection in a nonreacting supersonic combustor, J. Propuls. Power. 4 [197] D. Papamoschou, D.G. Hubbard, Visual observations of supersonic transverse jets,
(1988) 591–597, https://doi.org/10.2514/3.23105. Exp. Fluids. 14 (1993) 468–476, https://doi.org/10.1007/BF00190201.
[171] M. Lee, B.K. McMillin, J. Palmer, R.K. Hanson, Two-dimensional imaging of [198] D.E. Everett, M.A. Woodmansee, J.C. Dutton, M.J. Morris, Wall pressure
combustion phenomena in a shock tube using planar laser-induced fluorescence, measurements for a sonic jet injected transversely into a supersonic crossflow, J.
in: 29th Aerosp. Sci. Meet, 1991, https://doi.org/10.2514/6.1991-460. AIAA 91- Propuls. Power. 14 (1998) 861–868, https://doi.org/10.2514/2.5357.
0460, Reno, Nevada. [199] G.V. Candler, P.K. Subbareddy, N. Cymbalist, P. Dimotakis, Large-eddy
[172] M. Lee, B.K. McMillin, J. Palmer, R.K. Hanson, Planar fluorescence imaging of a simulation of autoignition-dominated supersonic combustion, in: 45th AIAA Fluid
transverse jet in a supersonic crossflow, J. Propuls. Power. 8 (1992) 729–735, Dyn. Conf, 2015, https://doi.org/10.2514/6.2015-3340. AIAA 2015-3340,
https://doi.org/10.2514/3.23542. Dallas, TX.
[173] A.D. Rothstein, P.J. Wantuck, A study of the normal injection of hydrogen into a [200] N. Cymbalist, P. Dimotakis, On autoignition-dominated supersonic combustion,
heated supersonicflow using planar laser-induced fluorescence, in: 28th Jt. in: 45th AIAA Fluid Dyn. Conf, 2015, https://doi.org/10.2514/6.2015-2315.
Propuls. Conf. Exhib, 1992, https://doi.org/10.2514/6.1992-3423. AIAA 92- AIAA 2015-2315, Dallas, TX.
3423, Nashville, TN. [201] D.W. Riggins, C.R. McClinton, Analysis of losses in supersonic mixing and
[174] M.G. Allen, T.E. Parker, W.G. Reinecke, H.H. Legner, R.R. Foutter, T.W. Rawlins, reacting flows, in: 27th Jt. Propuls. Conf, 1991, https://doi.org/10.2514/6.1991-
S.J. Davis, Fluorescence imaging of OH and NO in a model supersonic combustor, 2266. AIAA 91-2266, Sacramento, California.
AIAA J 31 (1993) 505–512, https://doi.org/10.2514/3.11358. [202] R.B. Mays, R.H. Thomas, J.A. Schetz, Low angle injection into a supersonic flow,
[175] T.E. Parker, M.G. Allen, R.R. Foutter, D.M. Sonnefroh, W.T. Rawlins, in: 25th Jt. Propuls. Conf, 1989, https://doi.org/10.2514/6.1989-2461. AIAA 89-
Measurements of OH and H2O for reacting flow in a supersonic combusting 2461, Monterey, CA.
ramjet combustor, J. Propuls. Power. 11 (1995) 1154–1161, https://doi.org/ [203] C.R. Mcclinton, The effect of injection angle on th e interaction between sonic
10.2514/3.23954. secondary jets and a supersonic free stream, Hampton, VA, 1972, https://ntrs.
[176] Q. Liu, A. Passaro, D. Baccarella, H. Do, Ethylene flame dynamics and inlet nasa.gov/citations/19720009559.
unstart in a model scramjet, J. Propuls. Power. 30 (2014) 1577–1585, https://doi. [204] J.A. Schetz, R.H. Thomas, F.S. Billig, Mixing of transverse jets and wall jets in
org/10.2514/1.B35214. supersonic flow, in: V. V Kozlov, A. V Dovgal (Eds.), Separated Flows Jets,
[177] Q. Liu, D. Baccarella, S. Hammack, T. Lee, C.D. Carter, H. Do, Influences of Springer Berlin Heidelberg, Berlin, Heidelberg, 1991, pp. 807–837.
freestream turbulence on flame dynamics in a supersonic combustor, AIAA J 55 [205] G.J. McCann, R.D.W. Bowersox, Experimental investigation of supersonic gaseous
(2016) 913–918, https://doi.org/10.2514/1.j055271. injection into a supersonic freestream, AIAA J 34 (1996) 317–323, https://doi.
[178] Q. Liu, D. Baccarella, T. Lee, S. Hammack, C.D. Carter, H. Do, Influences of inlet org/10.2514/3.13066.
geometry modification on scramjet flow and combustion dynamics, J. Propuls. [206] E.J. Fuller, R.H. Thomas, J.A. Schetz, Effects of yaw on low angle injection into a
Power. 33 (2017) 1179–1186, https://doi.org/10.2514/1.b36434. supersonic flow, in: 29th Aerosp. Sci. Meet, 1991, https://doi.org/10.2514/
[179] B.K. Mcmillin, J.M. Seitzman, R.K. Hanson, Comparison of NO and OH planar 6.1991-14. AIAA 91-0014, Reno, Nevada.
fluorescence temperature measurements in scramjet model flowfield, AIAA J 32 [207] E.J. Fuller, R.B. Mays, R.H. Thomas, J.A. Schetz, Mixing studies of helium in air at
(1994) 1945–1952, https://doi.org/10.2514/3.12237. high supersonic speeds, AIAA J 30 (1992) 2234–2243, https://doi.org/10.2514/
[180] W.M. VanLerberghe, J.G. Santiago, J.C. Dutton, R.P. Lucht, Mixing of a sonic 3.11210.
transverse jet injected into a supersonic flow, AIAA J 38 (2000) 470–479, https:// [208] J.P. Drummond, Enhancement of mixing and reaction in high-speed combustor
doi.org/10.2514/2.984. flowfields, in: Int. Colloq. Adv. Comput. Anal. Combust, 1997, pp. 1–14. Moscow,
[181] H. Takahashi, S. Ikegami, H. Oso, G. Masuya, M. Hirota, Quantitative imaging of Russia.
injectant mole fraction and density in a supersonic mixing, AIAA J 46 (2008) [209] J.M. Seiner, S.M. Dash, D.C. Kenzakowski, Historical survey on enhanced mixing
2935–2943, https://doi.org/10.2514/1.37783. in scramjet engines, J. Propuls. Power. 17 (2001) 1273–1286, https://doi.org/
[182] H. Takahashi, M. Hiroka, H. Oso, G. Masuya, Measurement of supersonic injection 10.2514/2.5876.
flowfield using acetone PLIF, Trans. Jpn. Soc. Aeronaut. Space Sci. 51 (2009) [210] D.W. Bogdanoff, Advanced injection and mixing techniques for scramjet
252–258, https://doi.org/10.2322/tjsass.51.252. combustors, J. Propuls. Power. 10 (1994) 183–190, https://doi.org/10.2514/
[183] V. Sick, High speed imaging in fundamental and applied combustion research, 3.23728.
Proc. Combust. Inst. 34 (2013) 3509–3530, https://doi.org/10.1016/j. [211] E.J. Gutmark, K.C. Schadow, K.H. Yu, Mixing enhancement in supersonic free
proci.2012.08.012. shear flows, Annu. Rev. Fluid Mech. 27 (1995) 375–417, https://doi.org/
[184] B. Thurow, N. Jiang, W. Lempert, Review of ultra-high repetition rate laser 10.1146/annurev.fl.27.010195.002111.
diagnostics for fluid dynamic measurements, Meas. Sci. Technol. 24 (2013), [212] P.E. Dimotakis, Turbulent mixing, Annu. Rev. Fluid Mech. 37 (2005) 329–356,
012002, https://doi.org/10.1088/0957-0233/24/1/012002. https://doi.org/10.1146/annurev.fluid.36.050802.122015.
[185] J.A. Boles, J.R. Edwards, R.A. Baurle, Large-Eddy/Reynolds-Averaged Navier- [213] D.M. Bushnell, Hypervelocity scramjet mixing enhancement, J. Propuls. Power.
Stokes simulations of sonic injection into mach 2 crossflow, AIAA J 48 (2010) 11 (1995) 1088–1090, https://doi.org/10.2514/3.51445.
1444–1456, https://doi.org/10.2514/1.J050066. [214] B.G. Northam, I. Greenberg, C.S. Byington, Evaluation of parallel injector
[186] S. Kawai, S.K. Lele, Large-eddy simulation of jet mixing in supersonic crossflows, configurations for supersonic combustion, in: 25th Jt. Propuls. Conf, 1989,
AIAA J 48 (2010) 2063–2083, https://doi.org/10.2514/1.J050282. https://doi.org/10.2514/6.1989-2525. AIAA 89-2525, Monterey, CA.
[187] Z.A. Rana, B. Thornber, D. Drikakis, Transverse jet injection into a supersonic [215] C.-M. Ho, E. Gutmark, Vortex induction and mass entrainment in a small-aspect-
turbulent cross-flow, Phys. Fluids. 23 (2011), 046103, https://doi.org/10.1063/ ratio elliptic jet, J. Fluid Mech. 179 (1987) 383–405, https://doi.org/10.1017/
1.3570692. S0022112087001587.
[188] H. Wang, Z. Wang, M. Sun, N. Qin, Hybrid Reynolds-averaged Navier-Stokes/ [216] A. Krothapalli, Y. Hsia, D. Baganoff, K. Karamcheti, The role of screech tones in
large-eddy simulation of jet mixing in a supersonic crossflow, Sci. China Technol. mixing of an underexpanded rectangular jet, J. Sound Vib. 106 (1986) 119–143,
Sci. 56 (2013) 1435–1448, https://doi.org/10.1007/s11431-013-5189-2. https://doi.org/10.1016/S0022-460X(86)80177-8.
[189] G.V. Candler, N. Cymbalist, P.E. Dimotakis, Wall-modeled large-eddy simulation [217] E. Gutmark, K.C. Schadow, C.J. Bicker, Near acoustic field and shock structure of
of autoignition-dominated supersonic combustion, AIAA J 55 (2017) 2410–2423, rectangular supersonic jets, AIAA J 28 (1990) 1163–1170, https://doi.org/
https://doi.org/10.2514/1.J055550. 10.2514/3.25187.
[190] M. Zhao, T. Zhou, T. Ye, M. Zhu, H. Zhang, Large eddy simulation of reacting flow [218] G. Raman, E. Rice, Instability modes excited by natural screech tones in a
in a hydrogen jet into supersonic cross-flow combustor with an inlet compression supersonic rectangular jets, in: 15th Aeroacoustics Conf, 1993, https://doi.org/
ramp, Int. J. Hydrogen Energy. 42 (2017) 16782–16792, https://doi.org/ 10.2514/6.1993-4321. AIAA 93-4321, Long Beach, CA.
10.1016/j.ijhydene.2017.04.250. [219] K.C. Schadow, K.J. Wilson, M.J. Lee, E. Gutmark, Enhancement of mixing in
[191] M. Sun, Y. Liu, Z. Hu, Turbulence decay in a supersonic boundary layer subjected reacting fuel-rich plumes issued from elliptical nozzles, J. Propuls. Power. 3
to a transverse sonic jet, J. Fluid Mech. 867 (2019) 216–249, https://doi.org/ (1987) 145–149, https://doi.org/10.2514/3.22966.
10.1017/jfm.2019.158. [220] E. Gutmark, K.C. Schadow, K.J. Wilson, Effect of convective Mach number on
[192] X. Chai, K. Mahesh, Simulations of high speed turbulent jets in crossflows, in: mixing of coaxial circular and rectangular jets, Phys. Fluids A Fluid Dyn. 3 (1991)
49th AIAA Aerosp. Sci. Meet. Incl. New Horizons Forum Aerosp. Expo, 2011, 29–36, https://doi.org/10.1063/1.857860.
https://doi.org/10.2514/6.2011-650. AIAA 2011-650, Orlando, Florida. [221] E. Gutmark, K.C. Schadow, K.J. Wilson, Subsonic and supersonic combustion
[193] F.P. Povinelli, L.A. Povinelli, Correlation of Secondary Sonic and Supersonic using noncircular injectors, J. Propuls. Power. 7 (1991) 240–249, https://doi.org/
Gaseous Jet Penetration into Supersonic Crossflows, 1971. Cleveland, Ohio. 10.2514/3.23317.
[194] Q. Liu, D. Baccarella, T. Lee, S.D. Hammack, C.D. Carter, H. Do, Investigation of [222] N.E. Masyakin, M.N. Polyanskii, On the possibility of blowing a gas jet into a
flow and combustion dynamics of an ethylene transverse jet in a model scramjet, supersonic flow without formation of a three-dimensional boundary-layer
in: 55th AIAA Aerosp. Sci. Meet, 2017, https://doi.org/10.2514/6.2017-0341. separation zone, Fluid Dyn 14 (1979) 459–461, https://doi.org/10.1007/
AIAA 2017-0341, Grapevine, Texas. BF01062456.
[195] M. Gamba, M.G. Mungal, R. Hanson, Ignition and near-wall burning in transverse [223] M.J. Barber, J.A. Schetz, L.A. Roe, Normal, sonic helium injection through a
hydrogen jets in supersonic crossflow, in: 49th AIAA Aerosp. Sci. Meet. Incl. New wedge-shaped orifice into supersonic flow, J. Propuls. Power. 13 (1997) 257–263,
Horizons Forum Aerosp. Expo, 2011, p. 2009, https://doi.org/10.2514/6.2011- https://doi.org/10.2514/2.5157.
319. AIAA 2011-319, Orlando, Florida. [224] R.W. Wlezien, V. Kibens, Influence of nozzle asymmetry on supersonic jets, AIAA
J 26 (1988) 27–33, https://doi.org/10.2514/3.9846.

72
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

[225] E.K. Longmire, J.K. Eaton, C.J. Elkins, Control of jet structure by crown-shaped [251] Y. Zhang, B. Wang, H. Zhang, S. Xue, Mixing enhancement of compressible planar
nozzle attachments, in: 29th Aerosp. Sci. Meet, 1991, https://doi.org/10.2514/ mixing layer impinged by oblique shock waves, J. Propuls. Power. 31 (2015)
6.1991-316. AIAA 91-0316, Reno, Nevada. 156–169, https://doi.org/10.2514/1.B35423.
[226] J. Bridges, C. Brown, Parametric testing of chevrons on single flow hot jets, in: [252] L. Zhang, J.Y. Choi, V. Yang, Supersonic combustion and flame stabilization of
10th AIAA/CEAS Aeroacoustics Conf, 2004, https://doi.org/10.2514/6.2004- coflow ethylene and air with splitter plate, J. Propuls. Power. 31 (2015)
2824. AIAA 2004-2824, Manchester, Great Britain. 1242–1255, https://doi.org/10.2514/1.B35740.
[227] M. Samimy, K. Zaman, M.F. Reeder, Effect of tabs on the flow and noise field of an [253] A. Kumar, D.M. Bushnell, M.Y. Hussaini, Mixing augmentation technique for
axisymmetric jet, AIAA J 31 (1993) 609–619, https://doi.org/10.2514/3.11594. hypervelocity scramjets, J. Propuls. Power. 5 (1989) 514–522, https://doi.org/
[228] K. Zaman, Spreading characteristics and thrust of jets from asymmetric nozzles, 10.2514/3.23184.
in: 34th Aerosp. Sci. Meet. Exhib, 1996, https://doi.org/10.2514/6.1996-200. [254] F.E. Marble, G.J. Hendricks, E.E. Zukoski, Progress toward shock enhancement of
AIAA 96-0200, Reno, NV. supersonic combustion processes, in: 23rd Jt. Propuls. Conf, 1987, https://doi.
[229] S. Tomioka, L.S. Jacobsen, J.A. Schetz, Sonic injection from diamond-shaped org/10.2514/6.1987-1880. AIAA 87-1880, San Diego, California.
orifices into a supersonic crossflow, J. Propuls. Power. 19 (2003) 104–114, [255] J.M. Budzinski, E.E. Zukoski, F.E. Marble, Rayleigh scattering measurements of
https://doi.org/10.2514/2.6086. shock enhanced mixing, in: 28th Jt. Propuls. Conf. Exhib, 1992, p. 3546, https://
[230] S. Tomioka, L. Jacobsen, J. Schetz, Interaction between a supersonic airstream doi.org/10.2514/6.1992-3546. Nashville, TN.
and a sonic jet injected through a diamond-shaped orifice, in: 38th Aerosp. Sci. [256] D.V. Gaitonde, Progress in shock wave/boundary layer interactions, Prog. Aerosp.
Meet. Exhib, 2000, https://doi.org/10.2514/6.2000-88. AIAA 2000-0088, Reno, Sci. 72 (2015) 80–99, https://doi.org/10.1016/j.paerosci.2014.09.002.
NV. [257] F.S. Billig, P.J. Waltrup, R.D. Stockbridge, Integral-rocket dual-combustion
[231] S. Tomioka, L. Jacobsen, J. Schetz, Angled injection through diamond-shaped ramjets: a new propulsion concept, J. Spacecr. Rockets. 17 (1980) 416–424,
orifices into a supersonic stream, in: 10th AIAA/NAL-NASDA-ISAS Int. Sp. Planes https://doi.org/10.2514/3.57760.
Hypersonic Syst. Technol. Conf, 2001, https://doi.org/10.2514/6.2001-1762. [258] S. Menon, Shock-wave-induced mixing enhancement in scramjet combustors, in:
AIAA 2001-1762, Kyoto, Japan. 27th Aerosp. Sci. Meet, 1989, https://doi.org/10.2514/6.1989-104. AIAA 89-
[232] R. Bowersox, H. Fan, D. Lee, Sonic injection into a mach 5.0 freestream through 0104, Reno, Nevada.
diamond orifices at various incidence, in: 39th AIAA/ASME/SAE/ASEE Jt. [259] H. Huh, J.F. Driscoll, Shock-wave-enhancement of the mixing and the stability
Propuls. Conf. Exhib, 2003, https://doi.org/10.2514/6.2003-5190. AIAA 2003- limits of supersonic hydrogen-air jet flames, Symp. Combust. 26 (1996)
5190, Huntsville, Alabama. 2933–2939, https://doi.org/10.1016/S0082-0784(96)80135-4.
[233] R.D.W. Bowersox, H. Fan, D. Lee, Sonic injection into a mach 5.0 freestream [260] X. Fang, C. Shen, M. Sun, Z. Hu, Effects of oblique shock waves on turbulent
through diamond orifices, J. Propuls. Power. 20 (2004) 280–287, https://doi.org/ structures and statistics of supersonic mixing layers, Phys. Fluids. 30 (2018)
10.2514/1.9254. 116101, https://doi.org/10.1063/1.5051015.
[234] R. Srinivasan, R.D.W. Bowersox, Simulation of transverse gaseous injection [261] C. Huete, A.L. Sánchez, F.A. Williams, J. Urzay, Diffusion-flame ignition by shock-
through diamond ports into supersonic freestream, J. Propuls. Power. 23 (2007) wave impingement on a supersonic mixing layer, J. Fluid Mech. 784 (2015)
772–782, https://doi.org/10.2514/1.18405. 74–108, https://doi.org/10.1017/jfm.2015.585.
[235] R. Srinivasan, R.D.W. Bowersox, Transverse injection through diamond and [262] C. Huete, A.L. Sánchez, F.A. Williams, Diffusion-flame ignition by shock-wave
circular ports into a mach 5.0 freestream, AIAA J 46 (2008) 1944–1962, https:// impingement on a hydrogen-air supersonic mixing layer, J. Propuls. Power. 33
doi.org/10.2514/1.29253. (2017) 256–263, https://doi.org/10.2514/1.B36236.
[236] K. Kobayashi, R.D.W. Bowersox, R. Srinivasan, C.D. Carter, K.-Y. Hsu, Flowfield [263] Y.R. Shau, D.S. Dolling, Experimental study of spreading rate enhancement of
studies of diamond-shaped fuel injector in a supersonic flow, J. Propuls. Power. high Mach numberturbulent shear layers, in: 25th Jt. Propuls. Conf, 1989,
23 (2007) 1168–1176, https://doi.org/10.2514/1.30000. https://doi.org/10.2514/6.1989-2458. AIAA 89-2458, Monterey, CA.
[237] P.M. Grossman, J.A. Schetz, L. Maddalena, Flush-wall, diamond-shaped fuel [264] N.T. Clemens, M.G. Mungal, Effects of sidewall disturbances on the supersonic
injector for high mach number scramjets, J. Propuls. Power. 24 (2008) 259–266, mixing layer, J. Propuls. Power. 8 (1992) 249–251, https://doi.org/10.2514/
https://doi.org/10.2514/1.29956. 3.23468.
[238] M.R. Gruber, A.S. Nejad, T.H. Chen, J.C. Dutton, Transverse injection from [265] P.J. Lu, K.C. Wu, On the shock enhancement of confined supersonic mixing flows,
circular and elliptic nozzles into a supersonic crossflow, J. Propuls. Power. 16 Phys. Fluids A Fluid Dyn. 3 (1991) 3046–3062, https://doi.org/10.1063/
(2000) 449–457, https://doi.org/10.2514/2.5609. 1.857849.
[239] G. Wang, L. Chen, X. Lu, Effects of the injector geometry on a sonic jet into a [266] J.-H. Kim, Y. Yoon, I.-S. Jeung, H. Huh, J.-Y. Choi, Numerical study of mixing
supersonic crossflow, Sci. China Physics, Mech. Astron. 56 (2013) 366–377, enhancement by shock waves in model scramjet engine, AIAA J 41 (2003)
https://doi.org/10.1007/s11433-012-4984-2. 1074–1080, https://doi.org/10.2514/2.2047.
[240] L. Foster, W. Engblom, Computation of transverse injection into supersonic [267] G. Park, C. Park, Y. Jin, H. Choi, J. Byun, K. Hwang, Ethylene transverse jets in
crossflow with various injector orifice geometries, in: 42nd AIAA Aerosp. Sci. supersonic crossflows, J. Propuls. Power. 31 (2015) 773–788, https://doi.org/
Meet. Exhib, 2004, https://doi.org/10.2514/6.2004-1199. AIAA 2004-1199, 10.2514/1.B35323.
Reno, Nevada. [268] M. Hagenmaier, J. Boles, R. Milligan, Scramjet Research with Flight-like Inflow
[241] H. Ogawa, Mixing characteristics of inclined fuel injection via various geometries Conditions, Wright-Patterson Air Force Base, OH, 2013. https://apps.dtic.mil/dt
for upstream-fuel-injected scramjets, J. Propuls. Power. 31 (2015) 1551–1566, ic/tr/fulltext/u2/a589252.pdf.
https://doi.org/10.2514/1.B35581. [269] T. Mai, Y. Sakimitsu, H. Nakamura, Y. Ogami, T. Kudo, H. Kobayashi, Effect of the
[242] H. Ogawa, Effects of injection angle and pressure on mixing performance of fuel incident shock wave interacting with transversal jet flow on the mixing and
injection via various geometries for upstream-fuel-injected scramjets, Acta combustion, Proc. Combust. Inst. 33 (2011) 2335–2342, https://doi.org/
Astronaut 128 (2016) 485–498, https://doi.org/10.1016/j. 10.1016/j.proci.2010.07.056.
actaastro.2016.08.008. [270] H. Nakamura, N. Sato, S. Ishida, Y. Ogami, H. Kobayashi, A study of interaction
[243] H. Ogawa, M. Kodera, Physical insight into fuel/air mixing with hypermixer between shock wave and cross-flow jet using particle tracking velocimetry, Trans.
injectors for scramjet engines, J. Propuls. Power. 31 (2015) 1423–1435, https:// Jpn. Soc. Aeronaut. Space Sci. 52 (2009) 81–88, https://doi.org/10.2322/
doi.org/10.2514/1.B35638. tjsass.52.81.
[244] K. Hirano, A. Matsuo, T. Kouchi, M. Izumikawa, S. Tomioka, New injector [271] W.M. Jungowski, Some self induced supersonic flow oscillations, Prog. Aerosp.
geometry for penetration enhancement of perpendicular jet into supersonic flow, Sci. 18 (1979) 151–175, https://doi.org/10.1016/0376-0421(77)90005-7.
in: 43rd AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2007, https://doi.org/ [272] K.J. Plotkin, Shock wave oscillation driven by turbulent boundary-layer
10.2514/6.2007-5028. AIAA 2007-5028, Cincinnati, OH. fluctuations, AIAA J 13 (1975) 1036–1040, https://doi.org/10.2514/3.60501.
[245] A.R. Hariharan, V. Babu, Transverse injection into a supersonic cross flow [273] H.H. Heller, D.G. Holmes, E.E. Covert, Flow-induced pressure oscillations in
through a circular injector with chevrons, J. Fluids Eng. 136 (2013), 021204, shallow cavities, J. Sound Vib. 18 (1971) 545–553, https://doi.org/10.1016/
https://doi.org/10.1115/1.4026018. 0022-460X(71)90105-2.
[246] E. Axdahl, A. Kumar, A. Wilhite, Study of forebody injection and mixing with [274] H. Heller, D. Bliss, The physical mechanism of flow-induced pressure fluctuations
application to hypervelocity airbreathing propulsion, in: 48th AIAA/ASME/SAE/ in cavities and concepts for their suppression, in: 2nd Aeroacoustics Conf, 1975,
ASEE Jt. Propuls. Conf. Exhib, 2012, https://doi.org/10.2514/6.2012-3924. https://doi.org/10.2514/6.1975-491. AIAA 75-491, Hampton, Virginia.
AIAA 2012-3924, Atlanta, Georgia. [275] M. Kawahashi, R. Bobone, E. Brocher, Oscillation modes in single-step Hartmann-
[247] T. Kouchi, G. Masuya, K. Hirano, A. Matsuo, S. Tomioka, Supersonic combustion Sprenger tubes, J. Acoust. Soc. Am. 75 (1984) 780–784, https://doi.org/10.1121/
using a stinger-shaped fuel injector, J. Propuls. Power. 29 (2013) 639–647, 1.390587.
https://doi.org/10.2514/1.B34524. [276] T. Kouchi, N. Sakuranaka, M. Izumikawa, S. Tomioka, Pulsed transverse injection
[248] H. Fan, R. Bowersox, Gaseous injection through diamond orifices at various applied to a supersonic flow, in: 43rd AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. &
incidence angles into a hypersonic freestream, in: 39th Aerosp. Sci. Meet. Exhib, Exhib., AIAA 2007-5405, Cincinnati, OH, https://doi.org/10.2514/6.2007-5405,
2001, https://doi.org/10.2514/6.2001-1050. AIAA 2001-1050, Reno, NV. 2007.
[249] T. Kouchi, S. Tomioka, K. Hirano, A. Matsuo, G. Masuya, Supersonic combustion [277] S.J. Kalidas, J. Kurian, Enhancement of supersonic mixing with the help of pulsed
using multiple stinger-shaped injectors, Trans. Jpn. Soc. Aeronaut. Space Sci. 60 injection, in: 43rd AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2007,
(2017) 56–59, https://doi.org/10.2322/tjsass.60.56. https://doi.org/10.2514/6.2007-5032. AIAA 2007-5032, Cincinnati, OH.
[250] D. Glawe, M. Samimy, S. lah, T. Chen, Effects of nozzle geometry on parallel [278] L.A. Smith, T. Ombrello, S. Okhovat, The effect of pulsed injection on the
injection from the base of an extended strut into a supersonic flow, in: 33rd entrainment into a cavity based flameholder in supersonic flow, in: 55th AIAA
Aerosp. Sci. Meet. Exhib, 1995, https://doi.org/10.2514/6.1995-522. AIAA Aerosp. Sci. Meet, 2017, https://doi.org/10.2514/6.2017-1538. AIAA 2017-
1995-0522, Reno, Nevada. 1538, Grapevine, Texas.

73
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

[279] H. Shi, G. Wang, X. Luo, J. Yang, X.-Y. Lu, Large-eddy simulation of a pulsed jet Propuls. Conf. Exhib, 2007, https://doi.org/10.2514/6.2007-5407. AIAA 2007-
into a supersonic crossflow, Comput. Fluids. 140 (2016) 320–333, https://doi. 5407, Cincinnati, OH.
org/10.1016/j.compfluid.2016.10.009. [306] N. Kubo, S. Tomioka, A. Murakami, K. Kudo, Mixing and combustion experiments
[280] A. Zheltovodov, E. Pimonov, D. Knight, Supersonic vortex breakdown control by with hyper-mixer injectors in a scramjet combustor, in: 50th AIAA/ASME/SAE/
energy deposition, in: 43rd AIAA Aerosp. Sci. Meet. Exhib, 2005, https://doi.org/ ASEE Jt. Propuls. Conf, 2014, https://doi.org/10.2514/6.2014-3873. AIAA 2014-
10.2514/6.2005-1048. AIAA 2005-1048, Reno, Nevada. 3873, Cleveland, OH.
[281] E. Lazar, G. Elliott, N. Glumac, Energy deposition applied to a transverse jet in a [307] N. Kubo, A. Murakami, K. Kudo, S. Tomioka, An experimental investigation on
supersonic crossflow, AIAA J 48 (2010) 1662–1672, https://doi.org/10.2514/1. combustion characteristics of hypermixer injectors–effects of the ‘swept’ applied
J050050. to hypermixer injector ramps, Procedia Eng 99 (2015) 954–960, https://doi.org/
[282] N. Gibbons, R. Gehre, S. Brieschenk, V. Wheatley, Blast wave-induced mixing in a 10.1016/j.proeng.2014.12.627.
laser ignited hypersonic flow, J. Fluids Eng. 140 (2017), 050902, https://doi.org/ [308] T. Sunami, P. Magre, A. Bresson, F. Grisch, M. Orain, M. Kodera, Experimental
10.1115/1.4038397. study of strut injectors in a supersonic combustor using OH-PLIF, in: AIAA/CIRA
[283] Y.M. Abul-Huda, M. Gamba, Enhanced combustion in supersonic flows using a 13th Int. Sp. Planes Hypersonics Syst. Technol. Conf, 2005, https://doi.org/
pulsed detonation, in: 54th AIAA Aerosp. Sci. Meet, 2016, https://doi.org/ 10.2514/6.2005-3304. AIAA 2005-3304, Capua, Italy.
10.2514/6.2016-1645. AIAA 2016-1645, San Diego, California. [309] T. Sunami, A. Murakami, K. Kudo, M. Kodera, M. Nishioka, Mixing and
[284] Y.M. Abul-huda, A Study of Combustion Augmentation in Supersonic Flows via a combustion control strategies for efficient scramjet operation in wide range of
Pulsed-Detonation Device, The University of Michigan, 2017. flight mach number, in: AIAA/AAAF 11th Int. Sp. Planes Hypersonic Syst.
[285] P.E. Dimotakis, Turbulent free shear layer mixing, in: 27th Aerosp. Sci. Meet, Technol. Conf, 2002, https://doi.org/10.2514/6.2002-5116. AIAA 2002-5116,
1989, https://doi.org/10.2514/6.1989-262. AIAA 89-0262, Reno, Nevada. Orleans, France.
[286] M.J. Gaston, A Study of Hypermixing Scramjet Fuel Injectors, Australian Defence [310] T. Sunami, F. Scheel, Analysis of mixing enhancement using streamwise vortices
Force Academy, 2002. in a supersonic combustor by application of laser diagnostics, in: AIAA/AAAF
[287] J. Swithenbank, I.W. Eames, S.B. Chin, B.C.R. Ewan, Z. Yang, J. Cao, X. Zhao, 11th Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2002, https://doi.org/
Turbulent mixing in supersonic combustion systems, in: E.T. Curran, S.N. 10.2514/6.2002-5203. AIAA 2002-5203, Orleans, France.
B. Murthy (Eds.), High-Speed Flight Propuls. Syst., American Institute of [311] T. Sunami, M. Wendt, M. Nishioka, Supersonic mixing and combustion control
Aeronautics and Astronautics, Washington DC, 1991, pp. 341–381, https://doi. using streamwise vortices, in: 34th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf.
org/10.2514/5.9781600866104.0341.0381. Exhib, 1998, https://doi.org/10.2514/6.1998-3271. AIAA 1998-3271, Cleveland,
[288] V.A. Vinogradov, Y.M. Shikhman, C. Segal, A review of fuel pre-injection in OH.
supersonic, chemically reacting flows, Appl. Mech. Rev. 60 (2007) 139, https:// [312] P. Gerlinger, P. Stoll, M. Kindler, F. Schneider, M. Aigner, Numerical investigation
doi.org/10.1115/1.2750346. of mixing and combustion enhancement in supersonic combustors by strut
[289] R.P. Fuller, P.-K. Wu, A.S. Nejad, J.A. Schetz, Comparison of physical and induced streamwise vorticity, Aerosp. Sci. Technol. 12 (2008) 159–168, https://
aerodynamic ramps as fuel injectors in supersonic flow, J. Propuls. Power. 14 doi.org/10.1016/j.ast.2007.04.003.
(1998) 135–145, https://doi.org/10.2514/2.5278. [313] J. Doster, P. King, M. Gruber, R. Maple, Pylon fuel injector design for a scramjet
[290] P.S. Kamath, C.R. Mcclinton, Scramjet combustor and nozzle computations, in: combustor, in: 43rd AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2007,
Proc. 18th Int. Counc. Aeronaut. Sci, 1992, pp. 703–711. Beijing, China. https://doi.org/10.2514/6.2007-5404. AIAA 2007-5404, Cincinnati, OH.
[291] R.J. Hartfield, S.D. Hollo, J.C. McDaniel, Experimental investigation of a [314] J.C. Doster, Hypermixer Pylon Fuel Injection for Scramjet Combustors, Air Force
supersonic swept ramp injector using laser-induced iodine fluorescence, in: 21st Institut of Technology, 2008. https://afit.idm.oclc.org/login?url=http://search.
Fluid Dyn. Plasma Dyn. Lasers Conf, 1990, https://doi.org/10.2514/6.1990- proquest.com/docview/304682133?accountid=26185%5Cnhttp://resolver.ebsc
1518. AIAA 90-1518, Seattle, WA. ohost.com/openurl?DOI=.
[292] R.J. Hartfield, S.D. Hollo, J.C. McDaniel, Experimental investigation of a [315] J.C. Doster, P.I. King, M.R. Gruber, C.D. Carter, M.D. Ryan, K.-Y. Hsu, In-stream
supersonic swept ramp injector using laser-induced iodine fluorescence, J. hypermixer fueling pylons in supersonic flow, J. Propuls. Power 25 (2009)
Propuls. Power. 10 (1994) 129–135, https://doi.org/10.2514/3.23721. 885–901, https://doi.org/10.2514/1.40179.
[293] R. Hönig, D. Theisen, R. Fink, R. Lachner, G. Kappler, D. Rist, P. Andresen, [316] F. Vergine, L. Maddalena, V. Miller, M. Gamba, Supersonic combustion and flame-
Experimental investigation of a SCRAMJET model combustor with injection holding characteristics of pylon injected hydrogen in a mach 2.4 high enthalpy
through a swept ramp using laser-induced fluorescence with tunable excimer flow, in: 50th AIAA Aerosp. Sci. Meet. Incl. New Horizons Forum Aerosp. Expo,
lasers, Symp. Combust. 26 (1996) 2949–2956, https://doi.org/10.1016/S0082- 2012, https://doi.org/10.2514/6.2012-333. AIAA 2012-0333, Nashville,
0784(96)80137-8. Tennessee.
[294] S.D. Stouffer, N.R. Baker, D.P. Capriotti, Effects of compression and expansion [317] F. Vergine, M. Crisanti, L. Maddalena, V. Miller, M. Gamba, Supersonic
ramp fuel injector configuration on scramjet combustion and heat transfer, in: combustion of pylon-injected hydrogen in high-enthalpy flow with imposed
31st Aerosp. Sci. Meet, 1993, https://doi.org/10.2514/6.1993-609. AIAA 93- vortex dynamics, J. Propuls. Power. 31 (2015) 89–103, https://doi.org/10.2514/
0609, Reno, NV. 1.B35330.
[295] C.-M. Hung, T. Barth, Computation of hypersonic flow through a narrow [318] C. Aguilera, A. Winkelmann, K. Yu, B. Pang, A. Ghosh, Supersonic mixing
expansion slot, AIAA J 28 (1990) 229–235, https://doi.org/10.2514/3.10379. enhancement and optimization using fin-guided fuel injection, in: 48th AIAA
[296] F.E. Marble, E.E. Zukoski, J.W. Jacobs, G.J. Hendricks, I.A. Waitz, Shock Aerosp. Sci. Meet. Incl. New Horizons Forum Aerosp. Expo, 2010, https://doi.
enhancement and control of hypersonic mixing and combustion, in: 26th Jt. org/10.2514/6.2010-1526. AIAA 2010-1526, Orlando, Florida.
Propuls. Conf, 1990, https://doi.org/10.2514/6.1990-1981. AIAA 90-1981, [319] C. Aguilera, B. Pang, A. Ghosh, A. Winkelmann, K. Yu, Scramjet mixing control
Orlando, FL. using fin-guided fuel injection, in: 45th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf.
[297] I.A. Waitz, F.E. Marble, Zuko, Vorticity generation by contoured wall injectors, in: Exhib, 2009, https://doi.org/10.2514/6.2009-5415. AIAA 2009-5415, Denver,
28th Jt. Propuls. Conf. Exhib, 1992, https://doi.org/10.2514/6.1992-3550. AIAA Colorado.
92-3550, Nashville, TN. [320] Y. Liu, M. Sun, C. Liang, J. Yu, G. Li, Flowfield structures of pylon-aided fuel
[298] I.A. Waitz, F.E. Marble, E.E. Zukoski, Investigation of a contoured wall injector for injection into a supersonic crossflow, Acta Astronaut 162 (2019) 306–313,
hypervelocity mixing augmentation, AIAA J 31 (1993) 1014–1021, https://doi. https://doi.org/10.1016/j.actaastro.2019.06.022.
org/10.2514/3.11723. [321] L. Li, W. Huang, L. Yan, Mixing augmentation induced by a vortex generator
[299] I.A. Waitz, E.M. Greitzer, C.S. Tan, Vortices in aero-propulsion systems, in: S. located upstream of the transverse gaseous jet in supersonic flows, Aerosp. Sci.
I. Green (Ed.), North-holl. Math. Libr., Springer, Dordrecht, 1995, pp. 471–531, Technol. 68 (2017) 77–89, https://doi.org/10.1016/j.ast.2017.05.016.
https://doi.org/10.1007/978-94-011-0249-0_11. [322] S.K. Cox, R.P. Fuller, J.A. Schetz, R.W. Walters, Vortical interactions generated by
[300] D.O. Davis, W.R. Hingst, Progress toward synergistic hypermixing nozzles, in: an injector array to enhance mixing in a supersonic flow, in: 32nd Aerosp. Sci.
27th Jt. Propuls. Conf, 1991, https://doi.org/10.2514/6.1991-2264. AIAA 91- Meet. Exhib, 1994, https://doi.org/10.2514/6.1994-708. AIAA 94-0708, Reno,
2264, Sacramento, California. NV.
[301] K. Itoh, Study on scramjet engine characteristics under mach 8 to 15 conditions, [323] S.-H. Lee, Characteristics of dual transverse injection in scramjet combustor, Part
in: D. Danesy (Ed.), Proc. 5th Eur. Symp. Aerothermodyn. Sp. Veh, 2004, p. 525. 1: mixing, J. Propuls. Power. 22 (2006) 1012–1019, https://doi.org/10.2514/
Cologne, Germany. 1.14180.
[302] M. Takahashi, K. Itoh, H. Tanno, T. Komuro, T. Sunami, K. Sato, S. Ueda, Study on [324] S.-H. Lee, Characteristics of dual transverse injection in scramjet combustor, Part
the high speed scramjet characteristics at Mach 10 to 15 flight condition, in: 2: combustion, J. Propuls. Power. 22 (2006) 1020–1026, https://doi.org/
Shock Waves, Springer Berlin Heidelberg, Berlin, Heidelberg, 2005, pp. 935–940, 10.2514/1.14185.
https://doi.org/10.1007/978-3-540-27009-6_142. [325] S.D. Hollo, J.C. McDaniel, R.J. Hartfield, Characterization of supersonic mixing in
[303] M. Takahashi, T. Sunami, H. Tanno, T. Komuro, M. Kodera, K. Itoh, Performance a nonreacting Mach 2 combustor, in: 30th Aerosp. Sci. Meet. Exhib, 1992, https://
characteristics of a scramjet engine at mach 10 to 15 flight condition, in: AIAA/ doi.org/10.2514/6.1992-93. AIAA 92-0093, Reno, NV.
CIRA 13th Int. Sp. Planes Hypersonics Syst. Technol. Conf, 2005, https://doi.org/ [326] S.D. Hollo, J.C. CDaniel, R.J. Artfield, Quantitative investigation of compressible
10.2514/6.2005-3350. AIAA 2005-3350, Capua, Italy. mixing - staged transverse injection into Mach 2 flow, AIAA J 32 (1994) 528–534,
[304] M. Takahashi, T. Komuro, K. Sato, M. Kodera, H. Tanno, K. Itoh, Performance https://doi.org/10.2514/3.12017.
characteristics of scramjet engine with different combustor shapes at [327] H. Takahashi, S. Ikegami, G. Masuya, M. Hirota, Extended quantitative
hypervelocity condition over mach 10 flight, in: 43rd AIAA/ASME/SAE/ASEE Jt. fluorescence imaging for multicomponent and staged injection into supersonic
Propuls. Conf. Exhib, 2007, https://doi.org/10.2514/6.2007-5395. AIAA 2007- crossflows, J. Propuls. Power. 26 (2010) 798–807, https://doi.org/10.2514/
5395, Cincinnati, OH. 1.47318.
[305] M. Kodera, V. Yang, M. Takahashi, K. Itoh, Ignition transient phenomena in a [328] H. Takahashi, G. Masuya, M. Hirota, Supersonic turbulent mixing structure in
scramjet engine at mach 12 flight condition, in: 43rd AIAA/ASME/SAE/ASEE Jt. staged injection flowfields, in: 48th AIAA Aerosp. Sci. Meet. Incl. New Horizons

74
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

Forum Aerosp. Expo, 2010, https://doi.org/10.2514/6.2010-753. AIAA 2010- [356] G. Binder, H. Didelle, Improvement of ejector thrust augmentation by pulsating or
753, Orlando, Florida. flapping jets, in: AGARD Conf. Proc. No. 308 Fluid Dyn. Jets with Appl. To V/
[329] W.O. Landsberg, V. Wheatley, A. Veeraragavan, Characteristics of cascaded fuel STOL, NATO, 1981, p. 22. Lisbon, Portugal.
injectors within an accelerating scramjet combustor, AIAA J 54 (2016) [357] G. Raman, M. Hailye, J.E. Rice, Flip-flop jet nozzle extended to supersonic flows,
3692–3700, https://doi.org/10.2514/1.J054815. AIAA J 31 (1993) 1028–1035, https://doi.org/10.2514/3.11725.
[330] A.S. Pudsey, R.R. Boyce, V. Wheatley, Hypersonic viscous drag reduction via [358] K.H. Yu, K.J. Wilson, K.C. Schadow, Effect of flame-holding cavities on
multiporthole injector arrays, J. Propuls. Power. 29 (2013) 1087–1096, https:// supersonic-combustion performance, J. Propuls. Power. 17 (2001) 1287–1295,
doi.org/10.2514/1.B34782. https://doi.org/10.2514/2.5877.
[331] M. Barzegar Gerdroodbary, D.D. Ganji, Y. Amini, Numerical study of shock wave [359] N. Sato, A. Imamura, S. Shiba, S. Takahashi, M. Tsue, M. Kono, Advanced mixing
interaction on transverse jets through multiport injector arrays in supersonic control in supersonic airstream with a wall-mounted cavity, J. Propuls. Power. 15
crossflow, Acta Astronaut 115 (2015) 422–433, https://doi.org/10.1016/j. (1999) 358–360, https://doi.org/10.2514/2.5433.
actaastro.2015.06.002. [360] K.H. Yu, K.C. Schadow, K.J. Kraeutle, E.J. Gutmark, Supersonic flow mixing and
[332] A. Anazadehsayed, M. Barzegar Gerdroodbary, Y. Amini, R. Moradi, Mixing combustion using ramp nozzle, J. Propuls. Power. 11 (1995) 1147–1153, https://
augmentation of transverse hydrogen jet by injection of micro air jets in doi.org/10.2514/3.23953.
supersonic crossflow, Acta Astronaut 137 (2017) 403–414, https://doi.org/ [361] K.H. Yu, R.A. Smith, K.J. Wilson, K.C. Schadow, Effect of excitation on supersonic
10.1016/j.actaastro.2017.05.007. jet afterburning, Combust. Sci. Technol. 113 (1996) 597–612, https://doi.org/
[333] S. Cox-Stouffer, M. Gruber, Effects of spanwise injection spacing on mixing 10.1080/00102209608935516.
characteristics of aerodynamic ramp injectors, in: 34th AIAA/ASME/SAE/ASEE [362] K.H. Yu, K.C. Schadow, Role of large coherent structures in turbulent
Jt. Propuls. Conf. Exhib, 1998, https://doi.org/10.2514/6.1998-3272. AIAA 98- compressible mixing, Exp. Therm. Fluid Sci. 14 (1997) 75–84, https://doi.org/
3272, Cleveland, OH. 10.1016/S0894-1777(96)00103-3.
[334] R. Fuller, P.-K. Wu, A. Nejad, J. Schetz, Fuel-vortex interactions for enhanced [363] V. Nenmeni, K. Yu, Cavity-induced mixing enhancement in confined supersonic
mixing in supersonic flow, in: 32nd Jt. Propuls. Conf. Exhib, 1996, https://doi. flows, in: 40th AIAA Aerosp. Sci. Meet. Exhib, 2002, https://doi.org/10.2514/
org/10.2514/6.1996-2661. AIAA 96-2661, Lake Buena Vista, FL. 6.2002-1010. AIAA 2002-1010, Reno, Nevada.
[335] S. Cox-Stouffer, M. Gruber, Further investigation of the effects of “aerodynamic [364] A. Ben-Yakar, R.K. Hanson, Cavity flame-holders for ignition and flame
ramp” design upon mixing characteristics, in: 35th Jt. Propuls. Conf. Exhib, 1999, stabilization in scramjets: an overview, J. Propuls. Power. 17 (2001) 869–877,
https://doi.org/10.2514/6.1999-2238. AIAA 99-2238, Los Angeles, California. https://doi.org/10.2514/2.5818.
[336] S. Cox-Stouffer, M. Gruber, Effects of injector yaw on mixing characteristics of [365] M.G. Owens, S. Tehranian, C. Segal, V.A. Vinogradov, Flame-Holding
aerodynamic ramp injectors, in: 37th Aerosp. Sci. Meet. Exhib, 1999, https://doi. configurations for kerosene combustion in a mach 1.8 airflow, J. Propuls. Power.
org/10.2514/6.1999-86. AIAA 99-0086, Reno, NV. 14 (1998) 456–461, https://doi.org/10.2514/2.5322.
[337] L.S. Jacobsen, An Integrated Aerodynamic-Ramp-Injector/Plasma-Torch-Igniter [366] W. Huang, M. Pourkashanian, L. Ma, D.B. Ingham, S. Bin Luo, Z.G. Wang,
for Supersonic Combustion Applications with Hydrocarbon Fuels, Virginia Investigation on the flameholding mechanisms in supersonic flows: backward-
Polytechnic Institute and State University, 2001. facing step and cavity flameholder, J. Vis. 14 (2011) 63–74, https://doi.org/
[338] L.S. Jacobsen, S.D. Gallimore, J.A. Schetz, W.F.O.’ Brien, Improved aerodynamic- 10.1007/s12650-010-0064-8.
ramp injector in supersonic flow, J. Propuls. Power. 19 (2003) 663–673, https:// [367] W. Huang, L. Jin, L. Yan, J. Tan, Influence of jet-to-crossflow pressure ratio on
doi.org/10.2514/2.6155. nonreacting and reacting processes in a scramjet combustor with backward-facing
[339] L. Jacobsen, J. Schetz, W. Ng, The flowfield near a multiport injector array in a steps, Int. J. Hydrogen Energy. 39 (2014) 21242–21250, https://doi.org/
supersonic flow, in: 34th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 1998, 10.1016/j.ijhydene.2014.10.073.
https://doi.org/10.2514/6.1998-3126. AIAA 98-3126, Cleveland, OH. [368] A. Thakur, C. Segal, Flameholding analyses in supersonic flow, in: 40th AIAA/
[340] K. Yu, E. Gutmark, R.A. Smith, K.C. Schadow, Supersonic jet excitation using ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2004, pp. 1–8, https://doi.org/
cavity-actuated forcing, in: 32nd Aerosp. Sci. Meet. Exhib, 1994, https://doi.org/ 10.2514/6.2004-3831. AIAA 2004-3831, Fort Lauderdale, Florida.
10.2514/6.1994-185. AIAA 94-0185, Reno, Nevada. [369] T. Niioka, H. Kobayashi, S. Hasegawa, K. Terada, Flame stabilization
[341] K.H. Yu, K.C. Schadow, Cavity-actuated supersonic mixing and combustion characteristics of strut divided into two parts in supersonic airflow, J. Propuls.
control, Combust. Flame. 99 (1994) 295–301, https://doi.org/10.1016/0010- Power. 11 (1995) 112–116, https://doi.org/10.2514/3.23847.
2180(94)90134-1. [370] G.B. Northam, C.A. Trexler, C.R. McClinton, Flameholding characteristics of a
[342] Y. Umeda, H. Maeda, R. Ishii, Discrete tones generated by the impingement of a swept-strut H2 fuel-injector for scramjet applications.pdf. 17th JANNAF Combust.
high-speed jet on a circular cylinder, Phys. Fluids. 30 (1987) 2380–2388, https:// Meet., Chemical Propulsion Information Agency, Hampton, Virginia, 1980.
doi.org/10.1063/1.866128. [371] R. Rogers, D. Capriotti, R. Guy, Experimental supersonic combustion research at
[343] W.C. Selerowicz, A.P. Szumowski, G.E.A. Meier, Self-excited compressible flow in NASA Langley, in: 20th AIAA Adv. Meas. Gr. Test. Technol. Conf, American
a pipe–collar nozzle, J. Fluid Mech. Digit. Arch. 228 (1991) 465, https://doi.org/ Institute of Aeronautics and Astronautics, Reston, Virigina, 1998, https://doi.org/
10.1017/S0022112091002781. 10.2514/6.1998-2506.
[344] M. Ponton, J.M. Seiner, The effects of initial jet exit conditions on plume [372] R.K. Soni, A. De, Investigation of strut-ramp injector in a Scramjet combustor:
resonance, in: 12th Aeroacoustic Conf, 1989, https://doi.org/10.2514/6.1989- effect of strut geometry, fuel and jet diameter on mixing characteristics, J. Mech.
1054. AIAA 89-1054, San Antonio, TX. Sci. Technol. 31 (2017) 1169–1179, https://doi.org/10.1007/s12206-017-0215-
[345] K. Yu, E. Gutmark, K.C. Schadow, Passive control of coherent vortices in 0.
compressible mixing layers, in: 3rd Shear Flow Conf, 1993, https://doi.org/ [373] T.A. Bovina, Studies of exchange between re-circulation zone behind the flame-
10.2514/6.1993-3262. AIAA 93-3262, Orlando, FL. holder and outer flow, Symp. Combust. 7 (1958) 692–696, https://doi.org/
[346] R.W. Wlezien, Nozzle geometry effects on supersonic jet interaction, AIAA J 27 10.1016/S0082-0784(58)80110-1.
(1989) 1361–1367, https://doi.org/10.2514/3.10272. [374] Y. Yoon, J. Donbar, J. Driscoll, Blowout and liftoff limits of a hydrogen jet flame
[347] C.K.W. Tam, The acoustic modes of a two-dimensional rectangular cavity, J. in a supersonic, heated, coflowing air stream, in: 31st Aerosp. Sci. Meet, 1993,
Sound Vib. 49 (1976) 353–364, https://doi.org/10.1016/0022-460X(76)90426- https://doi.org/10.2514/6.1993-446. AIAA 93-0446, Reno, Nevada.
0. [375] Y. Yoon, J.M. Donbar, J.F. Driscoll*, Blowout stability limits of a hydrogen jet
[348] C.K.W. Tam, P.J.W. Block, On the tones and pressure oscillations induced by flow flame in a supersonic, heated, coflowinq air stream, Combust. Sci. Technol. 97
over rectangular cavities, J. Fluid Mech. 89 (1978) 373–399, https://doi.org/ (1994) 137–156, https://doi.org/10.1080/00102209408935372.
10.1017/S0022112078002657. [376] R.W. Schefer, M. Namazian, J. Kelly, M. Perrin, Effect of confinement on bluff-
[349] C.K.W. Tam, Excitation of instability waves in a two-dimensional shear layer by body burner recirculation zone characteristics and flame stability, Combust. Sci.
sound, J. Fluid Mech. 89 (1978) 357–371, https://doi.org/10.1017/ Technol. 120 (1996) 185–211, https://doi.org/10.1080/00102209608935573.
S0022112078002645. [377] M. Owens, C. Segal, A.H. Auslender, Effects of mixing schemes on kerosene
[350] C.K.W. Tam, P.J. Morris, Tone excited jets, part V: a theoretical model and combustion in a supersonic airstream, J. Propuls. Power. 13 (1997) 525–531,
comparison with experiment, J. Sound Vib. 102 (1985) 119–151, https://doi.org/ https://doi.org/10.2514/2.5198.
10.1016/S0022-460X(85)80106-1. [378] Y. Zong, W. Bao, J. Chang, J. Hu, Q. Yang, J. Song, M. Wu, Effect of fuel injection
[351] C.K.W. Tam, F.Q. Hu, The instability and acoustic wave modes of supersonic Allocation on the combustion characteristics of a cavity-strut model scramjet, J.
mixing layers inside a rectangular channel, J. Fluid Mech. 203 (1989) 51–76, Aerosp. Eng. 28 (2015), 04014050, https://doi.org/10.1061/(ASCE)AS.1943-
https://doi.org/10.1017/S0022112089001370. 5525.0000374.
[352] C.K.W. Tam, F.Q. Hu, Resonant instability of ducted free supersonic mixing layers [379] K.-Y. Hsu, C.D. Carter, M.R. Gruber, T. Barhorst, S. Smith, Experimental study of
induced by periodic Mach waves, J. Fluid Mech. 229 (1991) 65, https://doi.org/ cavity-strut combustion in supersonic flow, J. Propuls. Power. 26 (2010)
10.1017/S002211209100294X. 1237–1246, https://doi.org/10.2514/1.45767.
[353] S. Martens, K. Kinzie, D. McLaughlin, Wave structure of coherent instabilities in a [380] J. Choi, C. Ghodke, S. Menon, Large-eddy simulation of cavity flame-holding in a
planar supersonic shear layer, in: 32nd Aerosp. Sci. Meet. Exhib, 1994, https:// mach 2.5 cross flow, in: 48th AIAA Aerosp. Sci. Meet. Incl. New Horizons Forum
doi.org/10.2514/6.1994-822. AIAA 94-0822, Reno, NV. Aerosp. Expo, 2010, https://doi.org/10.2514/6.2010-414. AIAA 2010-414,
[354] R.G. Adelgren, G.S. Elliott, J.B. Crawford, C.D. Carter, J.M. Donbar, D. Orlando, Florida.
F. Grosjean, Axisymmetric jet shear-layer excitation induced by laser energy and [381] C. Ghodke, J. Choi, S. Srinivasan, S. Menon, Large eddy simulation of supersonic
electric arc discharges, AIAA J 43 (2005) 776–791, https://doi.org/10.2514/ combustion in a cavity-strut flameholder, in: 49th AIAA Aerosp. Sci. Meet. Incl.
1.8548. New Horizons Forum Aerosp. Expo, 2011, https://doi.org/10.2514/6.2011-323.
[355] H. Viets, Flip-flop jet nozzle, AIAA J 13 (1975) 1375–1379, https://doi.org/ AIAA 2011-323, Orlando, Florida.
10.2514/3.60550. [382] G. Hongbin, C. Lihong, W. Dan, C. Xinyu, L. Zhi, Y. Xilong, Experimental
investigation on coupling characteristics of cavity flameholder and strut jet, in:

75
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

18th AIAA/3AF Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2012, https://doi. Meet, 2016, https://doi.org/10.2514/6.2016-0658. AIAA 2016-0658, San Diego,
org/10.2514/6.2012-5961. AIAA 2012-5961, Tours, France. California.
[383] C. Aguilera, K.H. Yu, Effect of fin-guided fuel injection on dual-mode scramjet [411] W. Xi, Z. Wang, M. Sun, W. Liu, Q. Li, Experimental investigation of ignition
operation, J. Propuls. Power. 33 (2017) 927–938, https://doi.org/10.2514/1. transient phase in model supersonic combustor, J. Aerosp. Eng. 27 (2014),
B36305. 04014009, https://doi.org/10.1061/(ASCE)AS.1943-5525.0000334.
[384] A. Oamjee, R. Sadanandan, Fuel injection location studies on pylon-cavity aided [412] T.X. Phuoc, Laser-induced spark ignition fundamental and applications, Opt.
jet in supersonic crossflow, Aerosp. Sci. Technol. 92 (2019) 869–880, https://doi. Lasers Eng. 44 (2006) 351–397, https://doi.org/10.1016/j.
org/10.1016/j.ast.2019.07.021. optlaseng.2005.03.008.
[385] A. Freeborn, P. King, M. Gruber, Leading edge pylon effects on a scramjet pylon- [413] S.A. O’Briant, S.B. Gupta, S.S. Vasu, Review: laser ignition for aerospace
cavity flameholder flowfield, in: 44th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. propulsion, Propuls. Power Res. 5 (2016) 1–21, https://doi.org/10.1016/j.
& Exhib, 2008, https://doi.org/10.2514/6.2008-4709. AIAA 2008-4709, jppr.2016.01.004.
Hartford, CT. [414] J. Zhang, J. Chang, J. Ma, C. Zhang, W. Bao, Investigation of flame establishment
[386] A.B. Freeborn, P.I. King, M.R. Gruber, Swept-leading-edge pylon effects on a and stabilization mechanism in a kerosene fueled supersonic combustor equipped
scramjet pylon-cavity flameholder flowfield, J. Propuls. Power. 25 (2009) with a thin strut, Aerosp. Sci. Technol. 70 (2017) 152–160, https://doi.org/
571–582, https://doi.org/10.2514/1.39546. 10.1016/j.ast.2017.08.005.
[387] K. Sathiyamoorthy, T.H. Danish, J. Srinivas, P. Manjunath, Experimental [415] J. Zhang, J. Chang, H. Tian, J. Li, W. Bao, Flame interaction characteristics in
investigation of supersonic combustion in a strut-cavity based combustor, Acta scramjet combustor equipped with strut/wall combined fuel injectors, Combust.
Astronaut 148 (2018) 285–293, https://doi.org/10.1016/j. Sci. Technol. 2202 (2019) 1–24, https://doi.org/10.1080/
actaastro.2018.05.014. 00102202.2019.1627342.
[388] G. Choubey, K.M. Pandey, Effect of different strut + wall injection techniques on [416] J. Hu, W. Bao, J. Chang, Flame transition in dual-mode scramjet combustor with
the performance of two-strut scramjet combustor, Int. J. Hydrogen Energy. 42 oxygen piloted ignition, J. Propuls. Power. 30 (2014) 1103–1107, https://doi.
(2017) 13259–13275, https://doi.org/10.1016/j.ijhydene.2017.04.024. org/10.2514/1.B35239.
[389] Q. Yang, W. Bao, Y. Zong, J. Chang, J. Hu, M. Wu, Combustion characteristics of a [417] W. Bao, J. Hu, Y. Zong, Q. Yang, M. Wu, J. Chang, D. Yu, Combustion
dual-mode scramjet injecting liquid kerosene by multiple struts, Proc. Inst. Mech. characteristic using O2-pilot strut in a liquid-kerosene-fueled strut-based dual-
Eng. Part G J. Aerosp. Eng. 229 (2015) 983–992, https://doi.org/10.1177/ mode scramjet, Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng. 227 (2013)
0954410014542447. 1870–1880, https://doi.org/10.1177/0954410012464455.
[390] M. Bouchez, X. Montazel, E. Dufour, Hydrocarbon fueled airbreathing propulsion [418] J. Hu, J. Qin, J. Chang, W. Bao, Y. Zong, Q. Yang, Combustion stabilization based
for high speed missiles, in: 34th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, on a center flame strut in a liquid kerosene fueled supersonic combustor, J.
1998, https://doi.org/10.2514/6.1998-3729. AIAA 1998-3729, Cleveland, OH. Therm. Sci. 22 (2013) 497–504, https://doi.org/10.1007/s11630-013-0654-6.
[391] S.-H. Zhu, X. Xu, Experimental study on flame transition in a two-stage struts [419] C.C. Rasmussen, J.F. Driscoll, C.D. Carter, K.-Y. Hsu, Characteristics of cavity-
dual-mode scramjet, J. Aerosp. Eng. 30 (2017), 06017002, https://doi.org/ stabilized flames in a supersonic flow, J. Propuls. Power. 21 (2005) 765–768,
10.1061/(ASCE)AS.1943-5525.0000740. https://doi.org/10.2514/1.15095.
[392] S. Zhu, X. Xu, P. Ji, Flame stabilization and propagation in dual-mode scramjet [420] C.C. Rasmussen, J.F. Driscoll, K.-Y. Hsu, J.M. Donbar, M.R. Gruber, C.D. Carter,
with staged-strut injectors, AIAA J 55 (2017) 171–179, https://doi.org/10.2514/ Stability limits of cavity-stabilized flames in supersonic flow, Proc. Combust. Inst.
1.J054974. 30 (2005) 2825–2833, https://doi.org/10.1016/j.proci.2004.08.185.
[393] S. Zhu, X. Xu, Q. Yang, Y. Jin, Intermittent back-flash phenomenon of supersonic [421] C.C. Rasmussen, S.K. Dhanuka, J.F. Driscoll, Visualization of flameholding
combustion in the staged-strut scramjet engine, Aerosp. Sci. Technol. 79 (2018) mechanisms in a supersonic combustor using PLIF, Proc. Combust. Inst. 31 (2007)
70–74, https://doi.org/10.1016/j.ast.2018.05.037. 2505–2512, https://doi.org/10.1016/j.proci.2006.08.007.
[394] K. Krishnamurty, Acoustic radiation from two-dimensional rectangular cutouts in [422] C. Rasmussen, J. Driscoll, Blowout limits of flames in high-speed airflows: critical
aerodynamic surfaces. https://doi.org/10.1515/9783110594843-153, 1955. damkohler number, in: 44th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib,
[395] H.E. Plumblee, J.S. Gibson, L.W. Lassiter, A Theoretical and Experimental 2008, https://doi.org/10.2514/6.2008-4571. AIAA 2008-4571, Reston, Virigina.
Investigation of the Acoustic Response of Cavities in an Aerodynamic Flow, 1962. [423] D. Micka, J. Driscoll, Reaction zone imaging in a dual-mode scramjet combustor
[396] X. Zhang, J.A. Edwards, An investigation of supersonic oscillatory cavity flows using CH-PLIF, in: 44th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2008,
driven by thick shear layers, Aeronaut. J. 94 (1990) 355–364, https://doi.org/ https://doi.org/10.2514/6.2008-5071. AIAA 2008-5071, Hartford, CT.
10.1017/S0001924000023319. [424] D.J. Micka, J.F. Driscoll, Combustion characteristics of a dual-mode scramjet
[397] J.E. Rossiter, Wind-Tunnel Experiments on the Flow over Rectangular Cavities at combustor with cavity flameholder, Proc. Combust. Inst. 32 II (2009) 2397–2404,
Subsonic and Transonic Speeds, 1964 internal-pdf://137.207.131.92/1966- https://doi.org/10.1016/j.proci.2008.06.192.
Rossiter-Wind-Tunnel Experiments on the F.pdf. [425] D. Micka, S. Torrez, J. Driscoll, Heat release distribution in a dual-mode scramjet
[398] H. Heller, J. Delfs, Cavity pressure oscillations: the generating mechanism combustor - measurements and modeling, in: 16th AIAA/DLR/DGLR Int. Sp.
visualized, J. Sound Vib. 196 (1996) 248–252, https://doi.org/10.1006/ Planes Hypersonic Syst. Technol. Conf, 2009, https://doi.org/10.2514/6.2009-
jsvi.1996.0480. 7362. AIAA 2009-7362, Bremen, Germany.
[399] A.J. Bilanin, E.E. Covert, Estimation of possible excitation frequencies for shallow [426] D.J. Micka, Combustion Stabilization, Structure, and Spreading in a Laboratory
rectangular cavities, AIAA J 11 (1973) 347–351, https://doi.org/10.2514/ Dual-Mode Scramjet Combustor, University of Michigan, 2010. http://search.
3.6747. proquest.com/docview/288155922?accountid=3611.
[400] Y.Y. Chan, Spatial waves in turbulent jets, Phys. Fluids. 17 (1974) 46, https://doi. [427] D.J. Micka, J.F. Driscoll, Stratified jet flames in a heated (1390K) air cross-flow
org/10.1063/1.1694612. with autoignition, Combust. Flame 159 (2012) 1205–1214, https://doi.org/
[401] Y.Y. Chan, Spatial waves of higher modes in an axisymmetric turbulent jet, Phys. 10.1016/j.combustflame.2011.10.013.
Fluids. 19 (1976) 2042, https://doi.org/10.1063/1.861404. [428] H. Wang, Z. Wang, M. Sun, H. Wu, Combustion modes of hydrogen jet combustion
[402] Y.Y. Chan, Wavelike eddies in a turbulent jet, AIAA J 15 (1977) 992–1001, in a cavity-based supersonic combustor, Int. J. Hydrogen Energy. 38 (2013)
https://doi.org/10.2514/3.60740. 12078–12089, https://doi.org/10.1016/j.ijhydene.2013.06.132.
[403] C.J. Moore, The role of shear-layer instability waves in jet exhaust noise, J. Fluid [429] H. Wang, Z. Wang, M. Sun, N. Qin, Combustion characteristics in a supersonic
Mech. 80 (1977) 321–367, https://doi.org/10.1017/S0022112077001700. combustor with hydrogen injection upstream of cavity flameholder, Proc.
[404] M.B. Sun, C. Gong, S.P. Zhang, J.H. Liang, W.D. Liu, Z.G. Wang, Spark ignition Combust. Inst. 34 (2013) 2073–2082, https://doi.org/10.1016/j.
process in a scramjet combustor fueled by hydrogen and equipped with multi- proci.2012.06.049.
cavities at Mach 4 flight condition, Exp. Therm. Fluid Sci. 43 (2012) 90–96, [430] J. Le, S. Yang, X. Wang, H. Li, Analysis and correlation of flame stability limits in
https://doi.org/10.1016/j.expthermflusci.2012.03.028. supersonic flow with cavity flameholder, in: 18th AIAA/3AF Int. Sp. Planes
[405] Y. Tian, S. Yang, J. Le, F. Zhong, X. Tian, Investigation of the effects of fuel Hypersonic Syst. Technol. Conf, 2012, https://doi.org/10.2514/6.2012-5948.
injector locations on ignition and flame stabilization in a kerosene fueled scramjet AIAA 2012-5948, Tours, France.
combustor, Aerosp. Sci. Technol. 70 (2017) 310–316, https://doi.org/10.1016/j. [431] S. O’Byrne, I. Stotz, A. Neely, R. Boyce, N. Mudford, F. Houwing, OH PLIF
ast.2017.08.012. imaging of supersonic combustion using cavity injection, in: AIAA/CIRA 13th Int.
[406] Z. Cai, J. Zhu, M. Sun, Z. Wang, Spark-enhanced ignition and flame stabilization Sp. Planes Hypersonics Syst. Technol. Conf, 2005, https://doi.org/10.2514/
in an ethylene-fueled scramjet combustor with a rear-wall-expansion geometry, 6.2005-3357. AIAA 2005-3357, Capua, Italy.
Exp. Therm. Fluid Sci. 92 (2018) 306–313, https://doi.org/10.1016/j. [432] E. Jeong, S. O’Byrne, I.-S. Jeung, A.F. Houwing, Supersonic combustion of
expthermflusci.2017.12.007. hydrogen fuel injection locations in a cavity-based combustor, in: 44th AIAA/
[407] T.M. Ombrello, C.D. Carter, C.-J. Tam, K.-Y. Hsu, Cavity ignition in supersonic ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2008, https://doi.org/10.2514/
flow by spark discharge and pulse detonation, Proc. Combust. Inst. 35 (2015) 6.2008-4576. AIAA 2008-4576, Hartford, CT.
2101–2108, https://doi.org/10.1016/j.proci.2014.07.068. [433] E. Jeong, I.-S. Jeung, S. O’Byrne, A.F.P. Houwing, Investigation of supersonic
[408] Z. Cai, J. Zhu, M. Sun, Z. Wang, X.-S. Bai, Laser-induced plasma ignition in a combustion with angled injection in a cavity-based combustor, J. Propuls. Power.
cavity-based scramjet combustor, AIAA J 56 (2018) 4884–4892, https://doi.org/ 24 (2008) 1258–1268, https://doi.org/10.2514/1.36519.
10.2514/1.J057081. [434] C. Liu, Z. Wang, M. Sun, H. Wang, P. Li, Characteristics of a cavity-stabilized
[409] Z. Cai, J. Zhu, M. Sun, Z. Wang, X.-S. Bai, Ignition processes and modes excited by hydrogen jet flame in a model scramjet combustor, AIAA J 57 (2019) 1624–1635,
laser-induced plasma in a cavity-based supersonic combustor, Appl. Energy. 228 https://doi.org/10.2514/1.J057346.
(2018) 1777–1782, https://doi.org/10.1016/j.apenergy.2018.07.079. [435] G. Masuya, T. Komuro, A. Murakami, N. Shinozaki, A. Nakamura, M. Murayamall,
[410] T. Ombrello, C.D. Carter, B. McGann, H. Do, D.M. Peterson, Establishing the K. Ohwaki, Ignition and combustion performance of scramjet combustors with
controlling parameters of ignition in high-speed flow, in: 54th AIAA Aerosp. Sci. fuel injection struts, J. Propuls. Power. 11 (1995) 301–307, https://doi.org/
10.2514/3.51425.

76
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

[436] T. Mitani, T. Kanda, T. Hiraiwa, Y. Igarashi, K. Nakahashi, Drags in scramjet [462] M. Berglund, C. Fureby, LES of supersonic combustion in a scramjet engine
engine testing: experimental and computational fluid dynamics studies, J. model, Proc. Combust. Inst. 31 (2007) 2497–2504, https://doi.org/10.1016/j.
Propuls. Power. 15 (1999) 578–583, https://doi.org/10.2514/2.5466. proci.2006.07.074.
[437] T. Mitani, N. Chinzei, T. Kanda, Reaction and mixing-controlled combustion in [463] F. Génin, S. Menon, Simulation of turbulent mixing behind a strut injector in
scramjet engines, J. Propuls. Power. 17 (2001) 308–314, https://doi.org/ supersonic flow, AIAA J 48 (2010) 526–539, https://doi.org/10.2514/1.43647.
10.2514/2.5743. [464] A. Mura, J.-F. Izard, Numerical simulation of supersonic nonpremixed turbulent
[438] S. Tomioka, A. Murakami, K. Kudo, T. Mitani, Combustion tests of a staged combustion in a scramjet combustor model, J. Propuls. Power. 26 (2010)
supersonic combustor with a strut, J. Propuls. Power. 17 (2001) 293–300, 858–868, https://doi.org/10.2514/1.48074.
https://doi.org/10.2514/2.5741. [465] C. Fureby, LES for supersonic combustion, in: 18th AIAA/3AF Int. Sp. Planes
[439] S. Tomioka, K. Kobayashi, K. Kudo, A. Murakami, T. Mitani, Effects of injection Hypersonic Syst. Technol. Conf, 2012, https://doi.org/10.2514/6.2012-5979.
configuration on performance of a staged supersonic combustor, J. Propuls. AIAA 2012-5979, Reston, Virigina.
Power. 19 (2003) 876–884, https://doi.org/10.2514/2.6178. [466] C. Fureby, E. Fedina, J. Tegnér, A computational study of supersonic combustion
[440] U. Wepler, W. Koschel, Numerical investigation of turbulent reacting flows in a behind a wedge-shaped flameholder, Shock Waves 24 (2014) 41–50, https://doi.
scramjet combustor model, 38th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. & org/10.1007/s00193-013-0459-2.
Exhib., AIAA 2002-3572, Reston, Virigina (2002), https://doi.org/10.2514/ [467] K. Wu, P. Zhang, W. Yao, X. Fan, LES study of flame stabilization in DLR hydrogen
6.2002-3572. supersonic combustor with strut injection, in: 21st AIAA Int. Sp. Planes
[441] E. Dufour, M. Bouchez, Computational analysis of a kerosene-fuelled scramjet, in: Hypersonics Technol. Conf, 2017, https://doi.org/10.2514/6.2017-2322. AIAA
10th AIAA/NAL-NASDA-ISAS Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2017-2322, Reston, Virginia.
2001, https://doi.org/10.2514/6.2001-1817. AIAA 2001-1817, Reston, Virigina. [468] C. Gong, M. Jangi, X.-S. Bai, J.-H. Liang, M.-B. Sun, Large eddy simulation of
[442] V.A. Vinogradov, S.A. Kobigsky, M.D. Petrov, Experimental investigation of hydrogen combustion in supersonic flows using an Eulerian stochastic fields
kerosene fuel combustion in supersonic flow, J. Propuls. Power. 11 (1995) method, Int. J. Hydrogen Energy. 42 (2017) 1264–1275, https://doi.org/
130–134, https://doi.org/10.2514/3.23850. 10.1016/j.ijhydene.2016.09.017.
[443] M. Dharavath, P. Manna, D. Chakraborty, Numerical exploration of mixing and [469] Z. Huang, G. He, S. Wang, F. Qin, X. Wei, L. Shi, Simulations of combustion
combustion in ethylene fueled scramjet combustor, Acta Astronaut 117 (2015) oscillation and flame dynamics in a strut-based supersonic combustor, Int. J.
305–318, https://doi.org/10.1016/j.actaastro.2015.08.014. Hydrogen Energy. 42 (2017) 8278–8287, https://doi.org/10.1016/j.
[444] Z. Huang, G. He, F. Qin, D. Cao, X. Wei, L. Shi, Large eddy simulation of ijhydene.2016.12.142.
combustion characteristics in a kerosene fueled rocket-based combined-cycle [470] A. Rajasekaran, V. Babu, Numerical simulation of three-dimensional reacting flow
engine combustor, Acta Astronaut 127 (2016) 326–334, https://doi.org/ in A model supersonic combustor, J. Propuls. Power. 22 (2006) 820–827, https://
10.1016/j.actaastro.2016.06.016. doi.org/10.2514/1.14952.
[445] Z. Shilong, F. Yuxin, Analysis of flow resistance and combustion characteristics in [471] M. Berglund, E. Fedina, C. Fureby, J. Tegnér, V. Sabel’nikov, Finite rate chemistry
the combined application of step and strut, Aerosp. Sci. Technol. 98 (2020) large-eddy simulation of self-ignition in supersonic combustion ramjet, AIAA J 48
105676, https://doi.org/10.1016/j.ast.2019.105676. (2010) 540–550, https://doi.org/10.2514/1.43746.
[446] N.R. Grady, R.W. Pitz, C.D. Carter, K.-Y. Hsu, C. Ghodke, S. Menon, Supersonic [472] C. Fureby, K. Nordin-Bates, K. Petterson, A. Bresson, V. Sabelnikov, A
flow over a ramped-wall cavity flame holder with an upstream strut, J. Propuls. computational study of supersonic combustion in strut injector and hypermixer
Power. 28 (2012) 982–990, https://doi.org/10.2514/1.B34394. flow fields, Proc. Combust. Inst. 35 (2015) 2127–2135, https://doi.org/10.1016/
[447] Y. Zhao, J. Liang, Y. Zhao, Non-reacting flow visualization of supersonic j.proci.2014.06.113.
combustor based on cavity and cavity–strut flameholder, Acta Astronaut 121 [473] K.-C. Lin, K. Jackson, R. Behdadnia, T.A. Jackson, F. Ma, V. Yang, Acoustic
(2016) 282–291, https://doi.org/10.1016/j.actaastro.2015.12.040. characterization of an ethylene-fueled scramjet combustor with a cavity
[448] D. Zhang, W. Song, Experimental study of cone-struts and cavity flameholders in a flameholder, J. Propuls. Power. 26 (2010) 1161–1170, https://doi.org/10.2514/
kerosene-fueled round scramjet combustor, Acta Astronaut 139 (2017) 24–33, 1.43338.
https://doi.org/10.1016/j.actaastro.2017.06.025. [474] Z. Wang, M. Sun, H. Wang, J. Yu, J. Liang, F. Zhuang, Mixing-related low
[449] J. Riehmer, E. Rabadan, A. Guelhan, B. Weigand, Experimental and numerical frequency oscillation of combustion in an ethylene-fueled supersonic combustor,
investigations of a scramjet model tested in the H2K blow down wind tunnel at Proc. Combust. Inst. 35 (2015) 2137–2144, https://doi.org/10.1016/j.
mach 7 flight condition, in: 19th AIAA Int. Sp. Planes Hypersonic Syst. Technol. proci.2014.09.005.
Conf, 2014, pp. 1–14, https://doi.org/10.2514/6.2014-2933. AIAA 2014-2933, [475] H. Wang, M. Sun, N. Qin, H. Wu, Z. Wang, Characteristics of oscillations in
Reston, Virginia. supersonic open cavity flows, flow, Turbul. Combust. 90 (2013) 121–142, https://
[450] J. Hu, J. Chang, W. Bao, Q. Yang, J. Wen, Experimental study of a flush wall doi.org/10.1007/s10494-012-9434-8.
scramjet combustor equipped with strut/wall fuel injection, Acta Astronaut 104 [476] H. Wang, Z. Wang, M. Sun, H. Wu, Nonlinear analysis of combustion oscillations
(2014) 84–90, https://doi.org/10.1016/j.actaastro.2014.07.012. in a cavity-based supersonic combustor, Sci. China Technol. Sci. 56 (2013)
[451] O.R. Kummitha, Numerical analysis of passive techniques for optimizing the 1093–1101, https://doi.org/10.1007/s11431-013-5198-1.
performance of scramjet combustor, Int. J. Hydrogen Energy. 42 (2017) [477] H. Wang, Z. Wang, M. Sun, Experimental study of oscillations in a scramjet
10455–10465, https://doi.org/10.1016/j.ijhydene.2017.01.148. combustor with cavity flameholders, Exp. Therm. Fluid Sci. 45 (2013) 259–263,
[452] W. Waidmann, F. Alff, M. Böhm, U. Brummund, W. Clauß, M. Oschwald, https://doi.org/10.1016/j.expthermflusci.2012.10.013.
Supersonic combustion of hydrogen/air in a SCRAMJET combustion chamber, Sp. [478] H. Wang, Z. Wang, M. Sun, N. Qin, Large-Eddy/Reynolds-averaged Navier–Stokes
Technol. 15 (1995) 421–429. simulation of combustion oscillations in a cavity-based supersonic combustor, Int.
[453] W. Waidmann, U. Brummund, J. Nuding, S.H. Chan, Experimental investigation J. Hydrogen Energy. 38 (2013) 5918–5927, https://doi.org/10.1016/j.
of supersonic ramjet combustion (SCRAMJET), in: 8th Transp. Phenom. Combust, ijhydene.2013.02.100.
Taylor & Francis, San Francisco, California, 1996, pp. 1473–1484. https://www. [479] M.L. Fotia, J.F. Driscoll, Ram-scram transition and flame/shock-train interactions
tib.eu/de/suchen/id/BLCP%3ACN018890368. in a model scramjet experiment, J. Propuls. Power. 29 (2013) 261–273, https://
[454] M. Oevermann, Numerical investigation of turbulent hydrogen combustion in a doi.org/10.2514/1.B34486.
SCRAMJET using flamelet modeling, Aerosp. Sci. Technol. 4 (2000) 463–480, [480] C. Aguilera, K.H. Yu, Scramjet to ramjet transition in a dual-mode combustor with
https://doi.org/10.1016/S1270-9638(00)01070-1. fin-guided injection, Proc. Combust. Inst. 36 (2017) 2911–2918, https://doi.org/
[455] F. Qin, Z. wei Huang, G. qiang He, S. Wang, X. geng Wei, B. Liu, Flame 10.1016/j.proci.2016.06.113.
stabilization mechanism study in a hydrogen-fueled model supersonic combustor [481] R. Yentsch, D. Gaitonde, Exploratory simulations of the HIFiRE 2 scramjet
under different air inflow conditions, Int. J. Hydrogen Energy. 42 (2017) flowpath, in: 48th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2012,
21360–21370, https://doi.org/10.1016/j.ijhydene.2017.06.237. https://doi.org/10.2514/6.2012-3772. AIAA 2012-3772, Atlanta, Georgia.
[456] K. Wu, P. Zhang, W. Yao, X. Fan, Numerical investigation on flame stabilization in [482] S. Yang, Y. Tian, J. Le, Flow oscillation in a scramjet combustor with air-
DLR hydrogen supersonic combustor with strut injection, Combust. Sci. Technol. throttling, in: 21st AIAA Int. Sp. Planes Hypersonics Technol. Conf, 2017, https://
189 (2017) 2154–2179, https://doi.org/10.1080/00102202.2017.1365847. doi.org/10.2514/6.2017-2301. AIAA 2017-2301, Xiamen, China.
[457] Y.-C. Cheng, K.-L. Pan, Simulation and design of A scramjet combustor enhanced [483] V. Yang, J. Li, J.Y. Choi, K.-C. Lin, Ignition transient in an ethylene fueled
by A porous cylinder burner, in: 2018 AIAA Aerosp. Sci. Meet, 2018, https://doi. scramjet engine with air throttling Part II: ignition and flame development, in:
org/10.2514/6.2018-0887. AIAA 2018-0887, Reston, Virginia. 48th AIAA Aerosp. Sci. Meet. Incl. New Horizons Forum Aerosp. Expo, 2010,
[458] B. Liu, G.-Q. He, F. Qin, J. An, S. Wang, L. Shi, Investigation of influence of https://doi.org/10.2514/6.2010-410. AIAA 2010-410, Orlando, Florida.
detailed chemical kinetics mechanisms for hydrogen on supersonic combustion [484] V. Yang, J. Li, J.Y. Choi, K.-C. Lin, Ignition transient in an ethylene fueled
using large eddy simulation, Int. J. Hydrogen Energy. 44 (2019) 5007–5019, scramjet engine with air throttling Part I: non-reacting flow development and
https://doi.org/10.1016/j.ijhydene.2019.01.005. mixing, in: 48th AIAA Aerosp. Sci. Meet. Incl. New Horizons Forum Aerosp. Expo,
[459] D. Zhang, D. Cao, G. He, B. Liu, F. Qin, Investigation on combustion mode and 2010, https://doi.org/10.2514/6.2010-409. AIAA 2010-409, Orlando, Florida.
heat release in a model scramjet engine affected by shocks, Int. J. Hydrogen [485] J. Li, L. Zhang, J.Y. Choi, V. Yang, K.-C. Lin, Ignition transients in a scramjet
Energy. 44 (2019) 28330–28341, https://doi.org/10.1016/j. engine with air throttling Part 1: nonreacting flow, J. Propuls. Power. 30 (2014)
ijhydene.2019.09.063. 438–448, https://doi.org/10.2514/1.B34763.
[460] Z. Huang, H. Zhang, Numerical investigations of mixed supersonic and subsonic [486] J. Li, F. Ma, V. Yang, K.-C. Lin, T. Jackson, A comprehensive study of ignition
combustion modes in a model combustor, Int. J. Hydrogen Energy. 45 (2020) transient in an ethylene-fueled scramjet combustor, in: 43rd AIAA/ASME/SAE/
1045–1060, https://doi.org/10.1016/j.ijhydene.2019.10.193. ASEE Jt. Propuls. Conf. Exhib, 2007, https://doi.org/10.2514/6.2007-5025.
[461] Z. Huang, G. He, F. Qin, X. Wei, Large eddy simulation of flame structure and AIAA 2007-5025, Cincinnati, OH.
combustion mode in a hydrogen fueled supersonic combustor, Int. J. Hydrogen [487] J. Noh, J.-Y. Choi, J.-R. Byun, J.-S. Lim, V. Yang, Numerical study of the auto-
Energy. 40 (2015) 9815–9824, https://doi.org/10.1016/j.ijhydene.2015.06.011. ignition of ethylene in an scramjet combustor with air throttling, in: 46th AIAA/

77
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2010, https://doi.org/10.2514/ nanosecond pulse duration plasma, Proc. Combust. Inst. 31 (2007) 3327–3334,
6.2010-7036. AIAA 2010-7036, Nashville, TN. https://doi.org/10.1016/j.proci.2006.07.126.
[488] Y. Tian, S. Yang, J. Le, Study on the effect of air throttling on flame stabilization [516] G. Lou, A. Bao, M. Nishihara, S. Keshav, Y. Utkin, I.V. Adamovich, Ignition in
of an ethylene fueled scramjet combustor, Int. J. Aerosp. Eng. 2015 (2015) 1–10, premixed hydrocarbon-air flows by repetitively pulsed, nanosecond pulse
https://doi.org/10.1155/2015/504684. duration plasma, in: 44th AIAA Aerosp. Sci. Meet. Exhib, 2006, https://doi.org/
[489] W. Deng, J. Le, S. Yang, W. Zhang, Y. Tian, Experimental research of air-throttling 10.2514/6.2006-1215. AIAA 2006-1215, Reno, NV.
ignition for a scramjet at Ma 6.5, Chinese J. Aeronaut. 30 (2017) 932–938, [517] T. Ombrello, Y. Ju, A. Fridman, Kinetic ignition enhancement of diffusion flames
https://doi.org/10.1016/j.cja.2017.03.017. by nonequilibrium magnetic gliding arc plasma, AIAA J 46 (2008) 2424–2433,
[490] Y. Tian, S. Yang, B. Xiao, J. Le, Experimental study on the effect of air throttling https://doi.org/10.2514/1.33005.
on supersonic combustion, Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng. 232 [518] T. Ombrello, Y. Ju, Kinetic ignition enhancement of H2 versus fuel-blended air
(2018) 472–480, https://doi.org/10.1177/0954410016680234. diffusion flames using nonequilibrium plasma, IEEE Trans. Plasma Sci. 36 (2008)
[491] F.E.C. Culick, T. Rogers, The response of normal shocks in diffusers, AIAA J 21 2924–2932, https://doi.org/10.1109/TPS.2008.2005987.
(1983) 1382–1390, https://doi.org/10.2514/3.60147. [519] N.L. Aleksandrov, S.V. Kindysheva, I.N. Kosarev, S.M. Starikovskaia, A.
[492] V. Yang, F.E.C. Culickt, Analysis of unsteady inviscid diffuser flow with a shock Y. Starikovskii, Mechanism of ignition by non-equilibrium plasma, Proc.
wave, J. Propuls. Power. 1 (1985) 222–228, https://doi.org/10.2514/3.22784. Combust. Inst. 32 (2009) 205–212, https://doi.org/10.1016/j.proci.2008.06.124.
[493] J.M. Donohue, Dual-mode scramjet flameholding operability measurements, J. [520] I.V. Adamovich, I. Choi, N. Jiang, J.-H. Kim, S. Keshav, W.R. Lempert,
Propuls. Power. 30 (2014) 592–603, https://doi.org/10.2514/1.B35016. E. Mintusov, M. Nishihara, M. Samimy, M. Uddi, Plasma assisted ignition and
[494] Y. Yang, Z. Wang, Y. Zhang, M. Sun, H. Wang, Flame stabilization with a rear- high-speed flow control: non-thermal and thermal effects, Plasma Sources Sci.
wall-expansion cavity in a supersonic combustor, Acta Astronaut 152 (2018) Technol. 18 (2009), 034018, https://doi.org/10.1088/0963-0252/18/3/034018.
752–756, https://doi.org/10.1016/j.actaastro.2018.09.027. [521] K.W. Hemawan, C.L. Romel, S. Zuo, I.S. Wichman, T.A. Grotjohn, J. Asmussen,
[495] K.-C. Lin, C.-J. Tam, I. Boxx, C. Carter, K. Jackson, M. Lindsey, Flame Microwave plasma-assisted premixed flame combustion, Appl. Phys. Lett. 89
characteristics and fuel entrainment inside a cavity flame holder of a scramjet (2006) 141501, https://doi.org/10.1063/1.2358213.
combustor, in: 43rd AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2007, [522] W. Kim, H. Do, M.G. Mungal, M.A. Cappelli, Plasma-discharge stabilization of jet
https://doi.org/10.2514/6.2007-5381. AIAA 2007-5381, Cincinnati, OH. diffusion flames, IEEE Trans. Plasma Sci. 34 (2006) 2545–2551, https://doi.org/
[496] K.-C. Lin, C.-J. Tam, K. Jackson, Study on the operability of cavity flameholders 10.1109/TPS.2006.886084.
inside a scramjet combustor, in: 45th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. [523] A. Vincent-Randonnier, S. Larigaldie, P. Magre, V. Sabel’nikov, Experimental
Exhib, 2009, https://doi.org/10.2514/6.2009-5028. AIAA 2009-5028, Denver, study of a methane diffusion flame under dielectric barrier discharge assistance,
Colorado. IEEE Trans. Plasma Sci. 35 (2007) 223–232, https://doi.org/10.1109/
[497] G. Retaureau, S. Menon, Experimental studies on flame stability of a fueled cavity TPS.2007.893249.
in a supersonic crossflow, in: 46th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. [524] A. Fridman, A. Gutsol, S. Gangoli, Y. Ju, T. Ombrello, Characteristics of gliding
Exhib, 2010, https://doi.org/10.2514/6.2010-6718. AIAA 2010-6718, Nashville, arc and its application in combustion enhancement, J. Propuls. Power. 24 (2008)
TN. 1216–1228, https://doi.org/10.2514/1.24795.
[498] C. Ghodke, G. Retaureau, J. Choi, S. Menon, Numerical and experimental studies [525] X. Yu, J. Peng, P. Yang, R. Sun, Y. Yi, Y. Zhao, D. Chen, J. Yu, Enhancement of a
of flame stability in a cavity stabilized hydrocarbon-fueled scramjet, in: 17th laminar premixed methane/oxygen/nitrogen flame speed using femtosecond-
AIAA Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2011, https://doi.org/ laser-induced plasma, Appl. Phys. Lett. 97 (2010), 011503, https://doi.org/
10.2514/6.2011-2365. AIAA 2011-2365, San Francisco, California. 10.1063/1.3457384.
[499] B.J. Tatman, R.D. Rockwell, C.P. Goyne, J.C. McDaniel, J.M. Donohue, [526] N. Chintala, A. Bao, G. Lou, I.V. Adamovich, Measurements of combustion
Experimental study of vitiation effects on flameholding in a cavity flameholder, efficiency in nonequilibrium RF plasma-ignited flows, Combust. Flame. 144
J. Propuls. Power. 29 (2013) 417–423, https://doi.org/10.2514/1.B34687. (2006) 744–756, https://doi.org/10.1016/j.combustflame.2005.08.040.
[500] D. Davis, R. Bowersox, Stirred reactor analysis of cavity flame holders for [527] H. Do, M.A. Cappelli, M.G. Mungal, Plasma assisted cavity flame ignition in
scramjets, in: 33rd Jt. Propuls. Conf. Exhib, 1997, https://doi.org/10.2514/ supersonic flows, Combust. Flame. 157 (2010) 1783–1794, https://doi.org/
6.1997-3274. AIAA 97-3274, Seattle, WA. 10.1016/j.combustflame.2010.03.009.
[501] D. Davis, R. Bowersox, Computational fluid dynamics analysis of cavity flame [528] H. Do, S. Im, M.A. Cappelli, M.G. Mungal, Plasma assisted flame ignition of
holders for scramjets, in: 33rd Jt. Propuls. Conf. Exhib, 1997, https://doi.org/ supersonic flows over a flat wall, Combust. Flame. 157 (2010) 2298–2305,
10.2514/6.1997-3270. AIAA 97-3270, Seattle, WA. https://doi.org/10.1016/j.combustflame.2010.07.006.
[502] K.-C. Lin, C.-J. Tam, K. Jackson, P. Kennedy, S. Williams, D. Olmstead, M. Collatz, [529] Y. Ju, W. Sun, Plasma assisted combustion: progress, challenges, and
Fueling study on scramjet operability enhancement, in: 45th AIAA/ASME/SAE/ opportunities, Combust. Flame. 162 (2015) 529–532, https://doi.org/10.1016/j.
ASEE Jt. Propuls. Conf. Exhib, 2009, https://doi.org/10.2514/6.2009-5116. combustflame.2015.01.017.
AIAA 2009-5116, Denver, Colorado. [530] A. Starikovskiy, N. Aleksandrov, Plasma-assisted ignition and combustion, Prog.
[503] V. Strokin, V. Grachov, The peculiarities of hydrogen combustion in model Energy Combust. Sci. 39 (2013) 61–110, https://doi.org/10.1016/j.
scramjet combustion, in: Proc. Int. Symp. Airbreathing Engines, 1997, pecs.2012.05.003.
pp. 374–384. ISABE 97-7056. [531] A. Starikovskiy, N. Aleksandrov, A. Rakitin, Plasma-assisted ignition and
[504] R.I. Ozawa, Survey of Basic Data on Flame Stabilization and Propagation for High deflagration-to-detonation transition, Philos. Trans. R. Soc. A Math. Phys. Eng.
Speed Combustion Systems, Marquardt Co., 1971. AFAPL-TR-70-81. Sci. 370 (2012) 740–773, https://doi.org/10.1098/rsta.2011.0344.
[505] Q. Liu, D. Baccarella, T. Lee, Influences of cavity on combustion stabilization in an [532] W. Sun, X. Gao, B. Wu, T. Ombrello, The effect of ozone addition on combustion:
axisymmetric scramjet, in: AIAA Scitech 2019 Forum, 2019, https://doi.org/ kinetics and dynamics, Prog. Energy Combust. Sci. 73 (2019) 1–25, https://doi.
10.2514/6.2019-1681. AIAA 2019-1681, San Diego, California. org/10.1016/j.pecs.2019.02.002.
[506] Q. Liu, D. Baccarella, B. McGann, T. Lee, Cavity-enhanced combustion stability in [533] W. Sun, Y. Ju, Nonequilibrium plasma-assisted combustion: a review of recent
an axisymmetric scramjet model, AIAA J (2019) 1–12, https://doi.org/10.2514/ progress, J. Plasma Fusion Res. 89 (2013) 208–219.
1.J058204. [534] S.M. Starikovskaia, Plasma assisted ignition and combustion, J. Phys. D. Appl.
[507] F.S. Billig, Design of supersonic combustors based on pressure-area fields, Symp. Phys. 39 (2006) R265–R299, https://doi.org/10.1088/0022-3727/39/16/R01.
Combust. (1967) 755–769, https://doi.org/10.1016/S0082-0784(67)80201-7. [535] S.M. Starikovskaia, Plasma-assisted ignition and combustion: nanosecond
[508] Q. Liu, D. Baccarella, B. McGann, T. Lee, Dual-mode operation and transition in discharges and development of kinetic mechanisms, J. Phys. D. Appl. Phys. 47
axisymmetric scramjets, AIAA J (2019) 1–14, https://doi.org/10.2514/1. (2014) 353001, https://doi.org/10.1088/0022-3727/47/35/353001.
J058391. [536] S. Leonov, Electrically driven supersonic combustion, Energies 11 (2018) 1733,
[509] D. Baccarella, Q. Liu, B.J. McGann, T. Lee, Combustion induced choking and https://doi.org/10.3390/en11071733.
unstart initiation in a circular constant-area supersonic flow, AIAA J (2019) 1–12, [537] P.L. Similon, R.N. Sudan, Plasma Turbulence, Annu. Rev. Fluid Mech. 22 (1990)
https://doi.org/10.2514/1.J057921. 317–347, https://doi.org/10.1146/annurev.fl.22.010190.001533.
[510] M. Gruber, S. Smith, T. Mathur, Experimental characterization of hydrocarbon- [538] J.S. Shang, R.L. Kimmel, J. Menart, S.T. Surzhikov, Hypersonic flow control using
fueled, axisymmetric, scramjet combustor flowpaths, in: 17th AIAA Int. Sp. Planes surface plasma actuator, J. Propuls. Power. 24 (2008) 923–934, https://doi.org/
Hypersonic Syst. Technol. Conf, 2011, https://doi.org/10.2514/6.2011-2311. 10.2514/1.24413.
AIAA 2011-2311, San Francisco, California. [539] A. Houpt, S.B. Leonov, F.H. Falempin, Control of hypersonic BL transition by
[511] C.S. Kalra, A.F. Gutsol, A.A. Fridman, Gliding arc discharges as a source of electrical discharge (feasibility study), in: 20th AIAA Int. Sp. Planes Hypersonic
intermediate plasma for methane partial oxidation, IEEE Trans. Plasma Sci. 33 Syst. Technol. Conf, 2015, https://doi.org/10.2514/6.2015-3602. AIAA 2015-
(2005) 32–41, https://doi.org/10.1109/TPS.2004.842321. 3602, Glasgow, Scotland.
[512] L. Bromberg, Plasma catalytic reforming of methane, Int. J. Hydrogen Energy. 24 [540] H. Yates, T.J. Juliano, E.H. Matlis, M.W. Tufts, Plasma-actuated flow control of
(1999) 1131–1137, https://doi.org/10.1016/S0360-3199(98)00178-5. hypersonic crossflow-induced boundary-layer transition in a mach-6 quiet tunnel,
[513] F.J. Weinberg, K. Hom, A.K. Oppenheim, K. Teichman, Ignition by plasma jet, in: 2018 AIAA Aerosp. Sci. Meet, 2018, https://doi.org/10.2514/6.2018-1076.
Nature 272 (1978) 341–343, https://doi.org/10.1038/272341a0. AIAA 2018-1076, Reston, Virginia.
[514] F. Wang, J.B. Liu, J. Sinibaldi, C. Brophy, A. Kuthi, C. Jiang, P. Ronney, M. [541] S.B. Leonov, D.A. Yarantsev, Near-surface electrical discharge in supersonic
A. Gundersen, Transient plasma ignition of quiescent and flowing air/fuel airflow: properties and flow control, J. Propuls. Power. 24 (2008) 1168–1181,
mixtures, IEEE Trans. Plasma Sci. 33 (2005) 844–849, https://doi.org/10.1109/ https://doi.org/10.2514/1.24585.
TPS.2005.845251. [542] F. Falempin, A.A. Firsov, D.A. Yarantsev, M.A. Goldfeld, K. Timofeev, S.
[515] G. Lou, A. Bao, M. Nishihara, S. Keshav, Y.G. Utkin, J.W. Rich, W.R. Lempert, I. B. Leonov, Plasma control of shock wave configuration in off-design mode of M =
V. Adamovich, Ignition of premixed hydrocarbon–air flows by repetitively pulsed, 2 inlet, Exp. Fluids. 56 (2015) 54, https://doi.org/10.1007/s00348-015-1928-4.

78
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

[543] H.-X. Huang, H.-J. Tan, S. Sun, Y.-C. Zhang, L. Cheng, Letter: transient interaction [570] A.L. Kuranov, E.G. Sheikin, Magnetohydrodynamic control on hypersonic aircraft
between plasma jet and supersonic compression ramp flow, Phys. Fluids. 30 under “Ajax” concept, J. Spacecr. Rockets. 40 (2003) 174–182, https://doi.org/
(2018), 041703, https://doi.org/10.1063/1.5028493. 10.2514/2.3951.
[544] A.A. Firsov, E. Dolgov, R. Rakhimov, M. Shurupov, S.B. Leonov, Mixing [571] D.W. Riggins, Analysis of the magnetohydrodynamic energy bypass engine for
enhancement by electrical discharge in supersonic airflow, in: 2018 AIAA Aerosp. high-speed airbreathing propulsion, J. Propuls. Power. 20 (2004) 779–792,
Sci. Meet, 2018, https://doi.org/10.2514/6.2018-1195. AIAA 2018-1195, https://doi.org/10.2514/1.4980.
Kissimmee, Florida. [572] A. Kuranov, A. Korabelnikov, Atmospheric cruise flight challenges for hypersonic
[545] S. Leonov, Y. Isaenkov, D. Yarantsev, M. Schneider, Fast mixing by pulse vehicles under the Ajax concept, J. Propuls. Power. 24 (2008) 1229–1247,
discharge in high-speed flow, in: 14th AIAA/AHI Sp. Planes Hypersonic Syst. https://doi.org/10.2514/1.24684.
Technol. Conf, 2006, https://doi.org/10.2514/6.2006-8129. AIAA 2006-8129, [573] R. Balasubramanian, K. Anandhanarayanan, R. Krishnamurthy, D. Chakraborty,
Canberra, Australia. Magnetohydrodynamic flow control of a hypersonic cruise vehicle based on AJAX
[546] S. Leonov, Y. Isaenkov, A. Firsov, Mixing intensification in high-speed flow by concept, J. Spacecr. Rockets. 53 (2016) 759–762, https://doi.org/10.2514/1.
unstable pulse discharge, in: 40th AIAA Plasmadynamics Lasers Conf, 2009, A33573.
https://doi.org/10.2514/6.2009-4074. AIAA 2009-4074, San Antonio, Texas. [574] I. Kimura, H. Aoki, M. Kato, The use of a plasma jet for flame stabilization and
[547] S. Im, H. Do, M.A. Cappelli, Dielectric barrier discharge control of a turbulent promotion of combustion in supersonic air flows, Combust. Flame. 42 (1981)
boundary layer in a supersonic flow, Appl. Phys. Lett. 97 (2010), 041503, https:// 297–305, https://doi.org/10.1016/0010-2180(81)90164-4.
doi.org/10.1063/1.3473820. [575] B.G. Northam, C.R. McClinton, T.C. Wagner, W.F. O’Brien, Development and
[548] S. Im, H. Do, G. Mungal, M. Cappelli, Experimental study and plasma control of an evaluation of a plasma jet flameholder for scramjets, in: 20th Jt. Propuls. Conf,
unstarting supersonic flow, in: 6th AIAA Flow Control Conf, 2012, https://doi. 1984, https://doi.org/10.2514/6.1984-1408. AIAA 84-1408, Cincinnati, Ohio.
org/10.2514/6.2012-2809. AIAA 2012-2809, New Orleans, Louisiana. [576] T.C. Wagner, W.F. O’Brien, B.G. Northam, J.M. Eggers, Plasma torch igniter for
[549] S. Im, H. Do, M.A. Cappelli, The manipulation of an unstarting supersonic flow by scramjets, J. Propuls. Power. 5 (1989) 548–554, https://doi.org/10.2514/
plasma actuator, J. Phys. D. Appl. Phys. 45 (2012) 485202, https://doi.org/ 3.23188.
10.1088/0022-3727/45/48/485202. [577] E. Barbi, J.R. Mahan, W.F. O’Brien, T.C. Wagner, Operating characteristics of a
[550] M.N. Shneider, S.O. Macheret, S.H. Zaidi, I.G. Girgis, R.B. Miles, Virtual shapes in hydrogen-argon plasma torch for supersonic combustion applications, J. Propuls.
supersonic flow control with energy addition, J. Propuls. Power. 24 (2008) Power. 5 (1989) 129–133, https://doi.org/10.2514/3.23126.
900–915, https://doi.org/10.2514/1.34136. [578] Y. Sato, M. Sayama, G. Masuya, T. Komuro, K. Kudou, A. Murakami, K. Tani,
[551] R.W. Ziemer, W.B. Bush, Magnetic field effects on bow shock stand-off distance, N. Chinzei, Experimental study on autoignition in a scramjet combustor, J.
Phys. Rev. Lett. 1 (1958) 58–59, https://doi.org/10.1103/PhysRevLett.1.58. Propuls. Power. 7 (1991) 657–658, https://doi.org/10.2514/3.23376.
[552] M.C. Smith, C.-S. Wu, Magnetohydrodynamic hypersonic viscous flow past a [579] S.B. Leonov, D.A. Yarantsev, A.P. Napartovich, I. V Kochetov, Plasma-assisted
blunt body, AIAA J 2 (1964) 963–965, https://doi.org/10.2514/3.2465. chemistry in high-speed flow, Plasma Sci. Technol. 9 (2007) 760–765, https://
[553] V. Bityurin, A. Klimov, S. Leonov, A. Bocharov, J. Lineberry, Assessment of a doi.org/10.1088/1009-0630/9/6/29.
concept of advanced flow/flight control for hypersonic flights in atmosphere, in: [580] K. Takita, Ignition and flame-holding by oxygen, nitrogen and argon plasma
9th Int. Sp. Planes Hypersonic Syst. Technol. Conf, 1999, https://doi.org/ torches in supersonic airflow, Combust. Flame. 128 (2002) 301–313, https://doi.
10.2514/6.1999-4820. AIAA 99-4820, Reston, Virigina. org/10.1016/S0010-2180(01)00354-6.
[554] J.-P. Petit, J. Geffray, F. David, MHD hypersonic flow control for aerospace [581] K. Takita, Ignition and flame-holding of H2 and CH4 in high temperature airflow
applications, in: 16th AIAA/DLR/DGLR Int. Sp. Planes Hypersonic Syst. Technol. by a plasma torch, Combust. Flame 132 (2003) 679–689, https://doi.org/
Conf, 2009, https://doi.org/10.2514/6.2009-7348. AIAA 2009-7348, Reston, 10.1016/S0010-2180(02)00518-7.
Virigina. [582] K. Takita, N. Abe, G. Masuya, Y. Ju, Ignition enhancement by addition of NO and
[555] T. Fujino, H. Sugita, M. Mizuno, I. Funaki, M. Ishikawa, Influences of electrical NO2 from a N2/O2 plasma torch in a supersonic flow, Proc. Combust. Inst. 31
conductivity of wall on magnetohydrodynamic control of Aeroynamic heating, J. (2007) 2489–2496, https://doi.org/10.1016/j.proci.2006.07.108.
Spacecr. Rockets. 43 (2006) 63–70, https://doi.org/10.2514/1.13770. [583] T. Kitagawa, A. Moriwaki, K. Murakami, K. Takita, G. Masuya, Ignition
[556] I.V. Adamovich, Plasma dynamics and flow control applications, in: Encycl. characteristics of methane and hydrogen using a plasma torch in supersonic flow,
Aerosp. Eng, John Wiley & Sons, Ltd, Chichester, UK, 2010, pp. 1–10, https://doi. J. Propuls. Power. 19 (2003) 853–858, https://doi.org/10.2514/2.6175.
org/10.1002/9780470686652.eae042. [584] N. Abe, R. Ohashi, K. Takita, G. Masuya, Y. Ju, Effects of NOx and HO2 on plasma
[557] A. Cristofolini, C.A. Borghi, G. Neretti, A. Schettino, E. Trifoni, F. Battista, ignition in a supersonic flow, in: 14th AIAA/AHI Sp. Planes Hypersonic Syst.
A. Passaro, D. Baccarella, Experimental investigations on the magneto-hydro- Technol. Conf, 2006, https://doi.org/10.2514/6.2006-7971. AIAA 2006-7971,
dynamic interaction around a blunt body in a hypersonic unseeded air flow, J. Canberra, Australia.
Appl. Phys. 112 (2012), 093304, https://doi.org/10.1063/1.4764105. [585] Y. Tan, C.G. Fotache, C.K. Law, Effects of NO on the ignition of hydrogen and
[558] A. Cristofolini, G. Neretti, C.A. Borghi, Plasma parameters and electromagnetic hydrocarbons by heated counterflowing air, Combust. Flame 119 (1999)
forces induced by the magneto hydro dynamic interaction in a hypersonic argon 346–355, https://doi.org/10.1016/S0010-2180(99)00064-4.
flow experiment, J. Appl. Phys. 112 (2012), 033302, https://doi.org/10.1063/ [586] Y. Sato, M. Sayama, K. Ohwaki, G. Masuya, T. Komuro, K. Kudou, A. Murakami,
1.4740052. K. Tani, Y. Wakamatsu, T. Kanda, N. Chinzei, I. Kimura, Effectiveness of plasma
[559] V. Fomichev, M. Yadrenkin, Strong MHD-intraction in hypersonic flows near torches for ignition and flameholding in scramjet, J. Propuls. Power. 8 (1992)
bodies, in: AIP Conf. Proc, 2017, 030020, https://doi.org/10.1063/1.5007478. 883–889, https://doi.org/10.2514/3.23565.
[560] V.P. Fomichev, M.A. Yadrenkin, Effect of electric discharges in magnetic field on [587] G. Masuya, K. Kudou, A. Murakami, K. Tani, T. Kanda, Y. Wakamatsu, N. Chinzei,
hypersonic flow around bodies, J. Phys. Conf. Ser. 1112 (2018), https://doi.org/ M. Sayama, K. Ohwaki, I. Kimura, Some governing parameters of plasma torch
10.1088/1742-6596/1112/1/012016. igniter/flameholder in a scramjet combustor, J. Propuls. Power. 9 (1993)
[561] T. Fujino, T. Yoshino, M. Ishikawa, Numerical analysis of reentry trajectory 176–181, https://doi.org/10.2514/3.23606.
coupled with magnetohydrodynamics flow control, J. Spacecr. Rockets 45 (2008) [588] S. Tomioka, T. Hiraiwa, N. Sakuranaka, A. Murakami, K. Sato, A. Matsui, Ignition
911–920, https://doi.org/10.2514/1.33385. strategy in a model scramjet, in: 32nd Jt. Propuls. Conf. Exhib, 1996, https://doi.
[562] S. Bobashev, N. Mende, V. Sakharov, D. Van Wie, MHD control of the separation org/10.2514/6.1996-3240. AIAA 96-3240, Lake Buena Vista, FL.
phenomenon in a supersonic xenon plasma flow, in: 41st Aerosp. Sci. Meet. Exhib, [589] K. Shuzenji, R. Kato, Two-stage plasma torch ignition in supersonic airflows, in:
2003, https://doi.org/10.2514/6.2003-168. AIAA 2003-0168, Reno, Nevada. 37th Jt. Propuls. Conf. Exhib, 2001, https://doi.org/10.2514/6.2001-3740. AIAA
[563] F. Cheng, X. Zhong, S. Gogineni, R.L. Kimmel, Magnetic-field effects on second- 2001-3740, Salt Lake City, Utah.
mode instability of a weakly ionized Mach 4.5 boundary layer, Phys. Fluids. 15 [590] Y. Matsubara, K. Takita, G. Masuya, Combustion enhancement in a supersonic
(2003) 2020–2040, https://doi.org/10.1063/1.1577565. flow by simultaneous operation of DBD and plasma jet, Proc. Combust. Inst. 34
[564] C.S. Kalra, M.N. Shneider, R.B. Miles, Numerical study of boundary layer (2013) 3287–3294, https://doi.org/10.1016/j.proci.2012.07.001.
separation control using magnetogasdynamic plasma actuators, Phys. Fluids. 21 [591] A. Bonanos, D. Sanders, J. Schetz, W. O’Brien, C. Goyne, R. Krauss, J. McDaniel,
(2009) 106101, https://doi.org/10.1063/1.3233658. Hot-flow testing of an integrated aero-ramp-injector/plasma-igniter for scramjets
[565] M. Nishihara, N. Jiang, J.W. Rich, W.R. Lempert, I.V. Adamovich, S. Gogineni, with hydrocarbon fuel, in: 12th AIAA Int. Sp. Planes Hypersonic Syst. Technol,
Low-temperature supersonic boundary layer control using repetitively pulsed 2003, https://doi.org/10.2514/6.2003-6987. AIAA 2003-6987, Norfolk, Virginia.
magnetohydrodynamic forcing, Phys. Fluids. 17 (2005) 106102, https://doi.org/ [592] A. Bonanos, J. Schetz, W. O’Brien, C. Goyne, Scramjet operability range studies of
10.1063/1.2084227. a multifuel integrated Aeroramp injector/plasma igniter, in: AIAA/CIRA 13th Int.
[566] S.O. Macheret, M.N. Shneider, R.B. Miles, Magnetohydrodynamic and Sp. Planes Hypersonics Syst. Technol. Conf, 2005, https://doi.org/10.2514/
Electrohydrodynamic control of hypersonic flows of weakly ionized plasmas, 6.2005-3425. AIAA 2005-3425, Capua, Italy.
AIAA J 42 (2004) 1378–1387, https://doi.org/10.2514/1.3971. [593] L.S. Jacobsen, C.D. Carter, R.A. Baurle, T.A. Jackson, S. Williams, D. Bivolaru,
[567] E. Gurijanov, P. Harsha, AJAX - new directions in hypersonic technology, in: Sp. S. Kuo, J. Barnett, C.-J. Tam, Plasma-assisted ignition in scramjets, J. Propuls.
Pl. Hypersonic Syst. Technol. Conf, 1996, https://doi.org/10.2514/6.1996-4609. Power. 24 (2008) 641–654, https://doi.org/10.2514/1.27358.
AIAA 1996-4609, Norfolk, VA. [594] N. Pavel, M. Bärwinkel, P. Heinz, D. Brüggemann, G. Dearden, G. Croitoru, O.
[568] R.J. Litchford, J.W. Cole, V.A. Bityurin, J.T. Lineberry, Thermodynamic cycle V. Grigore, Laser ignition - spark plug development and application in
analysis of magnetohydrodynamic-bypass hypersonic airbreathing engines, J. reciprocating engines, Prog. Quantum Electron. 58 (2018) 1–32, https://doi.org/
Propuls. Power. 17 (2001) 477–480, https://doi.org/10.2514/2.5769. 10.1016/j.pquantelec.2018.04.001.
[569] B.M. Burakhanov, A.P. Likhachev, S.A. Medin, V.A. Novikov, V.I. Okunev, V. [595] P.D. Ronney, Laser versus conventional ignition of flames, Opt. Eng. 33 (1994)
Y. Rickman, V.A. Zeigarnik, Advancement of scramjet magnetohydrodynamic 510–521, https://doi.org/10.1117/12.152237.
concept, J. Propuls. Power. 17 (2001) 1247–1252, https://doi.org/10.2514/
2.5871.

79
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

[596] S. Brieschenk, S. O’Byrne, H. Kleine, Laser-induced plasma ignition studies in a [621] S. Leonov, Y. Isaenkov, M. Shneider, A. Firsov, D. Yarsantsev, High-power
model scramjet engine, Combust. Flame. 160 (2013) 145–148, https://doi.org/ filamentary pulse discharge in supersonic flow, in: 48th AIAA Aerosp. Sci. Meet.
10.1016/j.combustflame.2012.08.011. Incl. New Horizons Forum Aerosp. Expo, 2010, https://doi.org/10.2514/6.2010-
[597] S. Brieschenk, H. Kleine, S. O’Byrne, Laser ignition of hypersonic air–hydrogen 259. AIAA 2010-259, Orlando, Florida.
flow, Shock Waves 23 (2013) 439–452, https://doi.org/10.1007/s00193-013- [622] S. Leonov, A. Firsov, D. Yarantsev, M. Bolshov, Y. Kuritsyn, V. Liger, V. Mironeko,
0447-6. Temperature measurement in plasma-assisted combustor by TDLAS, in: 43rd
[598] S. Brieschenk, S. O’Byrne, H. Kleine, Ignition characteristics of laser-ionized fuel AIAA Plasmadynamics Lasers Conf, 2012, https://doi.org/10.2514/6.2012-3181.
injected into a hypersonic crossflow, Combust. Flame. 161 (2014) 1015–1025, AIAA 2012-3181, New Orleans, Louisiana.
https://doi.org/10.1016/j.combustflame.2013.09.024. [623] S.B. Leonov, A. Vincent-Randonnier, V. Sabelnikov, A.A. Firsov, D. Yarantsev,
[599] G.S. Lee, Q. Liu, D. Baccarella, G.S. Elliott, T. Lee, A novel supersonic injection Plasma-assisted combustion in supersonic airflow: optimization of electrical
scheme for laser induced breakdown ignition, in: 2018 Aerodyn. Meas. Technol. discharge geometry, in: 52nd Aerosp. Sci. Meet, 2014, https://doi.org/10.2514/
Gr. Test. Conf, 2018, https://doi.org/10.2514/6.2018-4286. AIAA 2018-4286, 6.2014-0988. AIAA 2014-0988, National Harbor, Maryland.
Atlanta, Georgia. [624] S. Leonov, A. Firsov, D. Yarantsev, M. Bolshov, Y. Kuritsin, V. Liger,
[600] H. Do, C. Carter, Hydrocarbon fuel concentration measurement in reacting flows V. Mironenko, Dynamics of H2O temperature and concentration in zone of
using short-gated emission spectra of laser induced plasma, Combust. Flame. 160 plasma-assisted high-speed combustion, in: 49th AIAA Aerosp. Sci. Meet. Incl.
(2013) 601–609, https://doi.org/10.1016/J.COMBUSTFLAME.2012.12.002. New Horizons Forum Aerosp. Expo, 2011, https://doi.org/10.2514/6.2011-972.
[601] B. McGann, C.D. Carter, T.M. Ombrello, S. Hammack, T. Lee, H. Do, Gas property AIAA 2011-972, Orlando, Florida.
measurements in a supersonic combustor using nanosecond gated laser-induced [625] A.A. Firsov, S.B. Leonov, Numerical simulation of mixing enhancement by
breakdown spectroscopy with direct spectrum matching, Proc. Combust. Inst. 36 electrical discharge in supersonic airflow, in: AIAA Scitech 2019 Forum, 2019,
(2017) 2857–2864, https://doi.org/10.1016/j.proci.2016.06.089. https://doi.org/10.2514/6.2019-1349. AIAA 2019-1349, San Diego, California.
[602] B. McGann, T. Lee, T. Ombrello, C.D. Carter, S.D. Hammack, H. Do, Statistical [626] M.A. Shurupov, S.B. Leonov, A.A. Firsov, D.A. Yarantsev, Y.I. Isaenkov,
measurements of fuel mole fraction in scramjet cavity flameholder with a fuel Gasdynamic instabilities during decay of the submicrosecond spark discharge
surrogate, in: AIAA Scitech 2019 Forum, 2019, https://doi.org/10.2514/6.2019- channel, High Temp 52 (2014) 169–178, https://doi.org/10.1134/
1678. AIAA 2019-1678, San Diego, California. S0018151X14010192.
[603] L. Wermer, J.K. Lefkowitz, T. Ombrello, S. Im, Ignition enhancement by dual- [627] E. Tylczak, D. Peterson, G. Candler, Simulation of plasma-spark-enhanced mixing
pulse laser-induced spark ignition in a lean premixed methane-air flow, Proc. in jet in supersonic crossflow, in: 49th AIAA Aerosp. Sci. Meet. Incl. New Horizons
Combust. Inst. 37 (2019) 5605–5612, https://doi.org/10.1016/j. Forum Aerosp. Expo, 2011, https://doi.org/10.2514/6.2011-398. AIAA 2011-
proci.2018.06.025. 398, Orlando, Florida.
[604] X. Li, W. Liu, Y. Pan, L. Yang, B. An, Experimental investigation on laser-induced [628] J.R. McGuire, R.R. Boyce, N.R. Mudford, Radical-farm ignition processes in two-
plasma ignition of hydrocarbon fuel in scramjet engine at takeover flight dimensional supersonic combustion, J. Propuls. Power. 24 (2008) 1248–1257,
conditions, Acta Astronaut 138 (2017) 79–84, https://doi.org/10.1016/j. https://doi.org/10.2514/1.35562.
actaastro.2017.05.036. [629] J. Odam, A. Paull, Radical farming in scramjets. New Results Numer. Exp. Fluid
[605] L. Yang, B. An, X.P. Li, Y. Yu, X. Li, J. Liang, Q. Wang, Characterization of Mech. VI, Springer Berlin Heidelberg, Berlin, Heidelberg, 2007, pp. 276–283,
successive laser induced plasma ignition in an ethylene fuelled model scramjet https://doi.org/10.1007/978-3-540-74460-3_34.
engine, in: 21st AIAA Int. Sp. Planes Hypersonics Technol. Conf, 2017, https:// [630] R.R. Boyce, N.R. Mudford, J.R. McGuire, OH-PLIF visualisation of radical farming
doi.org/10.2514/6.2017-2353. AIAA 2017-2353, Xiamen, China. supersonic combustion flows, Shock Waves 22 (2012) 9–21, https://doi.org/
[606] B. An, Z. Wang, L. Yang, X. Li, J. Zhu, Experimental investigation on the impacts 10.1007/s00193-011-0346-7.
of ignition energy and position on ignition processes in supersonic flows by laser [631] P. Lorrain, S. Brieschenk, B. Capra, R. Boyce, A detailed investigation of
induced plasma, Acta Astronaut 137 (2017) 444–449, https://doi.org/10.1016/j. nominally 2-D radical-farming scramjet combustion, in: 18th AIAA/3AF Int. Sp.
actaastro.2017.05.013. Planes Hypersonic Syst. Technol. Conf, 2012, https://doi.org/10.2514/6.2012-
[607] X.-P. Li, W.-D. Liu, Y. Pan, S.-J. Liu, Investigation on ignition enhancement 5812. AIAA 2012-5812, Tours, France.
mechanism in a scramjet combustor with dual cavity, J. Propuls. Power. 32 [632] B.R. Capra, R.R. Boyce, S. Brieschenk, Numerical modeling of porous injection in
(2016) 439–447, https://doi.org/10.2514/1.B35880. a radical farming scramjet, in: 28th Int. Congr. Aeronaut. Sci, 2012. Brisbane,
[608] X. Li, L. Yang, J. Peng, X. Yu, J. Liang, R. Sun, Cavity ignition of liquid kerosene in QLD, Australia.
supersonic flow with a laser-induced plasma, Opt. Express. 24 (2016) 25362, [633] B. Capra, P. Lorrain, R. Boyce, S. Brieschenk, M. Kuhn, H. Hald, H2-O2 porous
https://doi.org/10.1364/OE.24.025362. fuel injection in a radical farming scramjet, in: 18th AIAA/3AF Int. Sp. Planes
[609] C. Dumitrache, M. Shneider, A.P. Yalin, Laser plasma formation in air using dual Hypersonic Syst. Technol. Conf, 2012, https://doi.org/10.2514/6.2012-5814.
pulse pre-ionization, in: 44th AIAA Plasmadynamics Lasers Conf, 2013, https:// AIAA 2012-5814, Tours, France.
doi.org/10.2514/6.2013-2632. AIAA 2013-2632, San Diego, CA. [634] B.R. Capra, R.R. Boyce, M. Kuhn, H. Hald, Porous versus porthole fuel injection in
[610] M.S. Bak, S. Im, M.A. Cappelli, Successive laser-induced breakdowns in a radical farming scramjet: numerical analysis, J. Propuls. Power. 31 (2015)
atmospheric pressure air and premixed ethane-air mixtures, Combust. Flame. 161 789–804, https://doi.org/10.2514/1.B35404.
(2014) 1744–1751, https://doi.org/10.1016/j.combustflame.2013.12.029. [635] J.C. Turner, M.K. Smart, Application of inlet injection to a three-dimensional
[611] X. Li, W. Liu, Y. Pan, L. Yang, B. An, J. Zhu, Characterization of ignition transient scramjet at mach 8, AIAA J 48 (2010) 829–838, https://doi.org/10.2514/1.
processes in kerosene-fueled model scramjet engine by dual-pulse laser-induced J050052.
plasma, Acta Astronaut 144 (2018) 23–29, https://doi.org/10.1016/j. [636] J.E. Barth, V. Wheatley, M.K. Smart, Effects of hydrogen fuel injection in a mach
actaastro.2017.12.018. 12 scramjet inlet, AIAA J 53 (2015) 2907–2919, https://doi.org/10.2514/1.
[612] E.A. Hassan, T. Ombrello, D.M. Peterson, Ignition and flame propagation in J053819.
cavity-fueled supersonic flameholder, in: AIAA Scitech 2019 Forum, 2019, [637] W.O. Landsberg, N.N. Gibbons, V. Wheatley, M.K. Smart, A. Veeraragavan, Flow
https://doi.org/10.2514/6.2019-1443. AIAA 2019-1443, San Diego, California. field manipulation via fuel injectors in scramjets, in: 21st AIAA Int. Sp. Planes
[613] T. Ombrello, X. Qin, Y. Ju, A. Gutsol, A. Fridman, Enhancement of combustion Hypersonics Technol. Conf, 2017, https://doi.org/10.2514/6.2017-2389. AIAA
and flame stabilization using stabilized non-equilibrium plasma, in: 43rd AIAA 2017-2389, Xiamen, China.
Aerosp. Sci. Meet. Exhib, 2005, https://doi.org/10.2514/6.2005-1194. AIAA [638] W.O. Landsberg, V. Wheatley, M.K. Smart, A. Veeraragavan, Performance of high
2005-1194, Reno, Nevada. mach number scramjets - tunnel vs flight, Acta Astronaut 146 (2018) 103–110,
[614] S.B. Leonov, D.A. Yarantsev, Plasma-induced ignition and plasma-assisted https://doi.org/10.1016/j.actaastro.2018.02.031.
combustion in high-speed flow, Plasma Sources Sci. Technol. 16 (2007) 132–138, [639] W.Y.K. Chan, S.A. Razzaqi, J.C. Turner, M.V. Suraweera, M.K. Smart, Freejet
https://doi.org/10.1088/0963-0252/16/1/018. testing of the HIFiRE 7 scramjet flowpath at mach 7.5, J. Propuls. Power. 34
[615] I.V. Adamovich, W.R. Lempert, J.W. Rich, Y.G. Utkin, M. Nishihara, Repetitively (2018) 844–853, https://doi.org/10.2514/1.b36652.
pulsed nonequilibrium plasmas for magnetohydrodynamic flow control and [640] A.D. Gardner, A. Paull, T.J. McIntyre, Upstream porthole injection in a 2-D
plasma-assisted combustion, J. Propuls. Power. 24 (2008) 1198–1215, https:// scramjet model, Shock Waves 11 (2002) 369–375, https://doi.org/10.1007/
doi.org/10.2514/1.24613. s001930200120.
[616] J.K. Lefkowitz, T. Ombrello, An exploration of inter-pulse coupling in nanosecond [641] E.W.K. Chang, S. Yang, G. Park, H. Choi, Ethylene flame-holding in double ramp
pulsed high frequency discharge ignition, Combust. Flame. 180 (2017) 136–147, flows, Aerosp. Sci. Technol. 80 (2018) 413–423, https://doi.org/10.1016/j.
https://doi.org/10.1016/j.combustflame.2017.02.032. ast.2018.07.012.
[617] S. Lovascio, T. Ombrello, J. Hayashi, S. Stepanyan, D. Xu, G.D. Stancu, C.O. Laux, [642] K. Kim, G. Park, S. Jin, Flameholding characteristics of ethylene-fueled model
Effects of pulsation frequency and energy deposition on ignition using scramjet in shock tunnel, Acta Astronaut 161 (2019) 446–464, https://doi.org/
nanosecond repetitively pulsed discharges, Proc. Combust. Inst. 36 (2017) 10.1016/j.actaastro.2019.02.022.
4079–4086, https://doi.org/10.1016/j.proci.2016.07.065. [643] R.J. Weber, J.S. MacKay, An Analysis of Ramjet Engines Using Supersonic
[618] H. Takana, I.V. Adamovich, H. Nishiyama, Computational simulation of Combustion, 1958. Cleveland, Ohio.
nanosecond pulsed discharge for plasma assisted ignition, J. Phys. Conf. Ser. 550 [644] R.A. Gross, W. Chinitz, A study of supersonic combustion, J. Aerosp. Sci. 27
(2014), 012051, https://doi.org/10.1088/1742-6596/550/1/012051. (1960) 517–524, https://doi.org/10.2514/8.8620.
[619] C.-C. Wang, Numerical simulation of combustion enhancement through a [645] J.A. Nicholls, Standing detonation waves, Symp, Combust 9 (1963) 488–498,
repetitive pulsed plasma actuator, J. Thermophys. Heat Transf. 30 (2016) https://doi.org/10.1016/S0082-0784(63)80059-4.
219–225, https://doi.org/10.2514/1.T4579. [646] R.P. Rhodes, P.M. Rubins, D.E. Chriss, Effect of heat release on flow parameters in
[620] S. Leonov, D. Yarantsev, Y. Isaenkov, Properties of filamentary electrical shock induced combustion, SAE Tech. Pap. (1964) 87–95, https://doi.org/
discharge in high-enthalpy flow, in: 43rd AIAA Aerosp. Sci. Meet. Exhib, 2005, 10.4271/640008.
https://doi.org/10.2514/6.2005-159. AIAA 2005-159, Reston, Virigina.

80
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

[647] P.M. Rubins, R.C. Bauer, Review of shock-induced supersonic combustion [676] J. Chan, J.P. Sislian, D. Alexander, Numerically simulated comparative
research and hypersonic applications, J. Propuls. Power. 10 (1994) 593–601, performance of a scramjet and shcramjet at mach 11, J. Propuls. Power. 26
https://doi.org/10.2514/3.23768. (2010) 1125–1134, https://doi.org/10.2514/1.48144.
[648] P.M. Rubins, R.P. Rhodes, Shock-induced combustion with oblique shocks- [677] A.P. Bruckner, The ram accelerator: overview and state of the art, in: Ram Accel,
comparison of experiment and kinetic calculations, AIAA J 1 (1963) 2778–2784, Springer Berlin Heidelberg, Berlin, Heidelberg, 1998, pp. 3–23, https://doi.org/
https://doi.org/10.2514/3.2172. 10.1007/978-3-642-46876-6_1.
[649] P.M. Rubins, T.H.M. Cunningham, Shock-induced combustion in a controlled area [678] G. Fusina, J.P. Sislian, B. Parent, formation and stability of near chapman-Jouguet
duct by means of oblique shocks, in: Aerosp. Sci. Meet, 1964, https://doi.org/ standing oblique detonation waves, AIAA J 43 (2005) 1591–1604, https://doi.
10.2514/6.1964-84. AIAA 1964-84, New York, NY. org/10.2514/1.9128.
[650] P.M. Rubins, T.H.M. Cunningham, Shock-induced supersonic combustion in a [679] J.M. Powers, K.A. Gonthier, Reaction zone structure for strong, weak overdriven,
constant-area duct, J. Spacecr. Rockets. 2 (1965) 199–205, https://doi.org/ and weak underdriven oblique detonations, Phys. Fluids A Fluid Dyn. 4 (1992)
10.2514/3.28151. 2082–2089, https://doi.org/10.1063/1.858378.
[651] P.M. Rubins, R.C. Bauer, A hypersonic ramjet analysis with premixed fuel [680] M.J. Grismer, J.M. Powers, Numerical predictions of oblique detonation stability
combustion, in: 2nd Propuls. Jt. Spec. Conf, 1966, https://doi.org/10.2514/ boundaries, Shock Waves 6 (1996) 147–156, https://doi.org/10.1007/
6.1966-648. AIAA 1966-648, Colorado Springs, CO. s001930050031.
[652] Y. Wang, J. Sislian, Numerical investigation of methane and air mixing in a [681] C. Li, K. Kailasanath, E.S. Oran, Detonation structures behind oblique shocks,
shcramjet inlet, in: 15th AIAA Int. Sp. Planes Hypersonic Syst. Technol. Conf, Phys. Fluids 6 (1994) 1600–1611, https://doi.org/10.1063/1.868273.
2008, https://doi.org/10.2514/6.2008-2533. AIAA 2008-2533, Dayton, Ohio. [682] C. Li, K. Kailasanath, E.S. Oran, Effects of boundary layers on oblique-detonation
[653] A. Hertzberg, A.P. Bruckner, D.W. Bogdanoff, Ram accelerator - a new chemical structures, in: 31st Aerosp. Sci. Meet, 1993, https://doi.org/10.2514/6.1993-450.
method for accelerating projectilesto ultrahigh velocities, AIAA J 26 (1988) AIAA 93-0450, Reno, NV.
195–203, https://doi.org/10.2514/3.9872. [683] Y. Fang, Z. Zhang, Z. Hu, Effects of boundary layer on wedge-induced oblique
[654] Kenneth K. Kuo, Principles of Combustion, second ed., John Wiley & Sons, Ltd, detonation structures in hydrogen-air mixtures, Int. J. Hydrogen Energy 44
Hoboken, New Jersey, 2005. https://www.wiley.com/en-us/Principles+of+Com (2019) 23429–23435, https://doi.org/10.1016/j.ijhydene.2019.07.005.
bustion%2C+2nd+Edition-p-9780471046899. [684] M. Yu, S. Miao, Initiation characteristics of wedge-induced oblique detonation
[655] F.S. Billig, External burning in supersonic streams, in: APL/JHU Rep, 1967. TG- waves in turbulence flows, Acta Astronaut 147 (2018) 195–204, https://doi.org/
912. 10.1016/j.actaastro.2018.04.022.
[656] F.S. Billig, External burning in supersonic streams, in: APL/JHU Tech. Dig, 1968, [685] P. Yang, H.D. Ng, H. Teng, Numerical study of wedge-induced oblique
pp. 2–10. detonations in unsteady flow, J. Fluid Mech. 876 (2019) 264–287, https://doi.
[657] D.L. Dugger, L. Monchick, External burning ramjets preliminary feasibility study, org/10.1017/jfm.2019.542.
in: APL/JHU Rep, 1967. TG-892. [686] A.A. Thaker, H.K. Chelliah, Numerical prediction of oblique detonation wave
[658] R.B. Morrison, Evaluation of the oblique detonation wave ramjet, in: s, 1978. Rep. structures using detailed and reduced reaction mechanisms, Combust. Theory
CR-145358. Model. 1 (1997) 347–376, https://doi.org/10.1080/713665338.
[659] R.B. Morrison, Oblique detonation wave ramjet, in: NASA Contract, 1980. Rep. [687] H.C. Yee, J.L. Shinn, Semi-implicit and fully implicit shock-capturing methods for
CR-159192. nonequilibrium flows, AIAA J 27 (1989) 299–307, https://doi.org/10.2514/
[660] G.P. Menees, H.G. Adelman, J.-L. Cambier, Analytical and experimental 3.10112.
validation of the oblique detonation wave engine concept, in: NASA Tech. Memo, [688] M.R. Kamel, C.I. Morris, I.G. Stouklov, R.K. Hanson, Plif imaging of hypersonic
1991. TM-102839. reactive flow around blunt bodies, Symp. Combust. 26 (1996) 2909–2915,
[661] H.G. Adelman, J.-L. Cambier, G.P. Menees, J.A. Balboni, Analytical and https://doi.org/10.1016/S0082-0784(96)80132-9.
experimental validation of the oblique detonation wave engine concept, in: 26th [689] C. Morris, M. Kamel, R. Hanson, Expansion tube investigation of ram-accelerator
Aerosp. Sci. Meet, 1988, https://doi.org/10.2514/6.1988-97. AIAA 88-0097, projectile flowfields, in: 32nd Jt. Propuls. Conf. Exhib, 1996, https://doi.org/
Reno, Nevada. 10.2514/6.1996-2680. AIAA 1996-2680, Lake Buena Vista, FL.
[662] J.-L. Cambier, H.G. Adelman, G.P. Menees, Numerical simulations of an oblique [690] C. Morris, M. Kamel, I. Stouklov, R. Hanson, PLIF imaging of the supersonic
detonation wave engine, J. Propuls. Power. 6 (1990) 315–323, https://doi.org/ reactive flows around projectiles in an expansion tube, in: 34th Aerosp. Sci. Meet.
10.2514/3.25436. Exhib, 1996, https://doi.org/10.2514/6.1996-855. AIAA 96-0855, Reno, NV.
[663] D.C. Alexander, J.P. Sislian, Computational study of the propulsive characteristics [691] C. Viguier, A. Gourara, D. Desbordes, Three-dimensional structure of stabilization
of a shcramjet engine, J. Propuls. Power. 24 (2008) 34–44, https://doi.org/ of oblique detonation wave in hypersonic flow, Symp, Combust 27 (1998)
10.2514/1.29951. 2207–2214, https://doi.org/10.1016/S0082-0784(98)80069-6.
[664] J. Sislian, J. Schumacher, A comparative study of hypersonic fuel/air mixing [692] C. Viguier, L.F.F. da Silva, D. Desbordes, B. Deshaies, Onset of oblique detonation
enhancement by ramp and cantilevered ramp injectors, in: 9th Int. Sp. Planes waves: comparison between experimental and numerical results for hydrogen-air
Hypersonic Syst. Technol. Conf, 1999, https://doi.org/10.2514/6.1999-4873. mixtures, Symp. Combust. 26 (1996) 3023–3031, https://doi.org/10.1016/
AIAA 99-4873, Norfolk, Virigina. S0082-0784(96)80146-9.
[665] B. Parent, J.P. Sislian, J. Schumacher, Numerical investigation of the turbulent [693] J. Verreault, A.J. Higgins, Initiation of detonation by conical projectiles, Proc.
mixing performance of a cantilevered ramp injector, AIAA J 40 (2002) Combust. Inst 33 (2011) 2311–2318, https://doi.org/10.1016/j.
1559–1566, https://doi.org/10.2514/2.1824. proci.2010.07.086.
[666] B. Parent, J.P. Sislian, Effect of geometrical parameters on the mixing [694] F.W. Ruegg, W.W. Dorsey, A missile technique for the study of detonation waves,
performance of cantilevered ramp injectors, AIAA J 41 (2003) 448–456, https:// J. Res. Natl. Bur. Stand. Sect. C Eng. Instrum. 66C (1962) 51, https://doi.org/
doi.org/10.2514/2.1966. 10.6028/jres.066C.007.
[667] J.P. Sislian, B. Parent, Hypervelocity fuel/air mixing in a shcramjet inlet, J. [695] H. Behrens, W. Struth, F. Wecken, Studies of hypervelocity firings into mixtures of
Propuls. Power. 20 (2004) 263–272, https://doi.org/10.2514/1.9252. hydrogen with air or with oxygen, Symp. Combust. 10 (1965) 245–252, https://
[668] T.E. Schwartzentruber, J.P. Sislian, B. Parent, Suppression of premature ignition doi.org/10.1016/S0082-0784(65)80169-2.
in the premixed inlet flow of a shcramjet, J. Propuls. Power. 21 (2005) 87–94, [696] T.-Y. Toong, Instabilities in reacting flows, Acta Astronaut 1 (1974) 317–344,
https://doi.org/10.2514/1.7003. https://doi.org/10.1016/0094-5765(74)90101-5.
[669] D.C. Alexander, J.P. Sislian, B. Parent, Hypervelocity fuel/air mixing in mixed- [697] A. Matsuo, T. Fujiwara, K. Fujii, Flow features of shock-induced combustion
compression inlets of shcramjets, AIAA J 44 (2006) 2145–2155, https://doi.org/ around projectile travelling at hypervelocities, in: 31st Aerosp. Sci. Meet, 1993,
10.2514/1.12630. https://doi.org/10.2514/6.1993-451. AIAA 93-0451, Reno, Nevada.
[670] J.P. Sislian, R.P. Martens, T.E. Schwartzentruber, B. Parent, Numerical simulation [698] S. Maeda, R. Inada, J. Kasahara, A. Matsuo, Visualization of the non-steady state
of a real shcramjet flowfield, J. Propuls. Power. 22 (2006) 1039–1048, https:// oblique detonation wave phenomena around hypersonic spherical projectile,
doi.org/10.2514/1.14895. Proc. Combust. Inst. 33 (2011) 2343–2349, https://doi.org/10.1016/j.
[671] J.P. Sislian, R. Dudebout, J. Schumacher, M. Islam, T. Redford, Incomplete mixing proci.2010.06.066.
and off-design effects on shock-induced combustion ramjet performance, J. [699] S. Maeda, J. Kasahara, A. Matsuo, Oblique detonation wave stability around a
Propuls. Power. 16 (2000) 41–48, https://doi.org/10.2514/2.5529. spherical projectile by a high time resolution optical observation, Combust. Flame
[672] R. Veraar, A. Mayer, J. Verreault, R. Stowe, R. Farinaccio, P. Harris, Proof-of- 159 (2012) 887–896, https://doi.org/10.1016/j.combustflame.2011.09.001.
Principle experiment of a shock-induced combustion ramjet, in: 16th AIAA/DLR/ [700] S. Maeda, S. Sumiya, J. Kasahara, A. Matsuo, Initiation and sustaining
DGLR Int. Sp. Planes Hypersonic Syst. Technol. Conf, 2009, https://doi.org/ mechanisms of stabilized Oblique Detonation Waves around projectiles, Proc.
10.2514/6.2009-7432. AIAA 2009-7432, Reston, Virigina. Combust. Inst 34 (2013) 1973–1980, https://doi.org/10.1016/j.
[673] S.A. Ashford, G. Emanuel, Oblique detonation wave engine performance proci.2012.05.035.
prediction, J. Propuls. Power. 12 (1996) 322–327, https://doi.org/10.2514/ [701] S. Maeda, S. Sumiya, J. Kasahara, A. Matsuo, Scale effect of spherical projectiles
3.24031. for stabilization of oblique detonation waves, Shock Waves 25 (2015) 141–150,
[674] D. Couture, A. DeChamplain, R. Stowe, P. Harris, W. Halswijk, J.-L. Moerel, https://doi.org/10.1007/s00193-015-0549-4.
Comparison of scramjet and shcramjet propulsion for an hypersonic waverider [702] S. Yungster, S. Eberhardt, A.P. Bruckner, Numerical simulation of hypervelocity
configuration, in: 44th AIAA/ASME/SAE/ASEE Jt. Propuls. Conf. Exhib, 2008, projectiles in detonable gases, AIAA J 29 (1991) 187–199, https://doi.org/
https://doi.org/10.2514/6.2008-5171. AIAA 2008-5171, Hartford, CT. 10.2514/3.10564.
[675] M.K.L. O’Neill, M.J. Lewis, Design tradeoffs on scramjet engine integrated [703] J.K. Ahuja, S.N. Tiwari, D.J. Singh, Investigation of hypersonic shock-induced
hypersonic waverider vehicles, J. Aircr. 30 (1993) 943–952, https://doi.org/ combustion in a hydrogen-air system, in: 30th Aerosp. Sci. Meet. Exhib, 1992,
10.2514/3.46438. https://doi.org/10.2514/6.1992-339. AIAA 92-0339, Reno, Nevada.

81
Q. Liu et al. Progress in Aerospace Sciences 119 (2020) 100636

[704] D.J. Singh, J.K. Ahuja, M.H. Carpenter, Numerical simulation of shock-induced [708] Y. Kamiyama, A. Matsuo, Flow features of shock-induced combustion around
combustion/detonation, Comput. Syst. Eng. 3 (1992) 201–215, https://doi.org/ cylindrical projectiles, Proc. Combust. Inst. 28 (2000) 671–677, https://doi.org/
10.1016/0956-0521(92)90106-S. 10.1016/S0082-0784(00)80268-4.
[705] J.K. Ahuja, S.N. Tiwari, D.J. Singh, Hypersonic shock-induced combustion in a [709] P.R. Ess, J.P. Sislian, C.B. Allen, Blunt-body generated detonation in viscous
hydrogen-air system, AIAA J 33 (1995) 173–176, https://doi.org/10.2514/ hypersonic ducted flows, J. Propuls. Power. 21 (2005) 667–680, https://doi.org/
3.12354. 10.2514/1.5579.
[706] A. Matsuo, K. Fujii, Detailed mechanism of the unsteady combustion around [710] J. Kasahara, T. Fujiwara, T. Endo, Chapman-jouguet oblique detonation structure
hypersonic projectiles, AIAA J 34 (1996) 2082–2089, https://doi.org/10.2514/ around hypersonic projectiles, AIAA J 39 (2001) 1553–1561, https://doi.org/
3.13355. 10.2514/2.1480.
[707] Y. Ju, G. Masuya, A. Sasoh, Numerical and theoretical studies on detonation [711] J. Verreault, Initiation of Gaseous Detonation by Conical Projectiles, McGill
initiation by a supersonic projectile, Symp. Combust. 27 (1998) 2225–2231, University, 2011.
https://doi.org/10.1016/S0082-0784(98)80071-4.

82

You might also like