Sei sulla pagina 1di 20

Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Progress in Aerospace Sciences


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

Supersonic spray combustion subject to scramjets: Progress and challenges


Zhaoxin Rena, Bing Wangb,∗, Gaoming Xiangb, Dan Zhaoc, Longxi Zhenga
a
School of Power and Energy, Northwestern Polytechnical University, Xi'an, 710072, China
b
School of Aerospace Engineering, Tsinghua University, Beijing, 100084, China
c
Department of Mechanical Engineering, College of Engineering, University of Canterbury, Private Bag 4800, Christchurch, 8140, New Zealand

A R T I C LE I N FO A B S T R A C T

Keywords: Supersonic spray combustion is one of the most significant physiochemical processes that occurs in the scramjet
upersonic combustion propulsion system, and related research can motivate the development of scramjet engines. This paper reviews
Spray combustion the research on supersonic spray combustion that has been conducted in the past few decades and focuses on the
Atomization key physiochemical processes and associated fluid physical mechanisms. Supersonic spray combustion involves
Shock induced combustion
not only the typical processes during the combustion of liquid fuel, such as atomization, dispersion, evaporation,
Detonation
Scramjet engines
mixing and ignition, but also complicated interactions among the spray, turbulence, shock waves and chemical
reactions in supersonic flows. The interactions commonly affect the combustion performance in terms of igni-
tion, stability and efficiency. The present work outlines the current research status for spray combustion in
subsonic flows. Then a brief description of the basic features of supersonic flow and combustion is provided. For
the supersonic combustor in the scramjet engine, the research results and challenges regarding supersonic flow
and spray in mixing layers and jets are discussed. The stabilization of supersonic spray combustion and the
control methods are introduced and summarized. An outline of the effects of shock waves on combustion, and
the shock-induced deflagration and detonation is provided as an overview of the research and development
progress since shock waves typically occur in supersonic combustors. Finally, the potential challenges and issues
that are encountered in the fundamental research, including numerical models and approaches and experimental
techniques and databases, and the applications of supersonic spray combustion are highlighted.

1. Introduction consists of three parts: an inlet, a combustion chamber and a nozzle.


Hence, this engine has a simple geometric configuration and lacks the
Supersonic/hypersonic flight technologies have become points of complex rotating parts that are used in the turbine engine. In contrast to
interest in the field of international aerospace technology. The devel- the ramjet engine, there are no physical throats in the pure scramjet
opment and application of supersonic/hypersonic vehicles are of high since the streams have supersonic speed in the internal passage. The
strategic significance to the next generation of atmospheric transpor- concept of supersonic combustion was successfully demonstrated for
tation [1]. The supersonic air-breathing engine is a key component of the first time under laboratory conditions in the 1960s [5]. Although
hypersonic vehicles. When the flight Mach number of the aircraft ex- the prototype engine was tested in a wind tunnel and the corresponding
ceeds five, to avoid aerodynamic loss and excessive dissociation of combustion features were measured, flight verification was not com-
oxygen and nitrogen in the airstream, the airflow in the combustion pleted until the 1970s, when the Langley research center of NASA
chamber should be maintained at the supersonic speed [2]. Fuel is in- launched a series of research projects with the objective of developing
jected into the supersonic airflow in the combustor of the supersonic the integrated fuselage of scramjet engine and designing a hypersonic
air-breathing engine, and mixes with the local air, thereby resulting in a vehicle with a flight Mach number of 7. The Central Institute of Avia-
supersonic combustion. The associated combustion process is com- tion Motors (CIAM) designed a dual-mode scramjet engine with hy-
pleted in milliseconds. drogen fuel and performed its first flight test in 1991. After that, re-
The concept of the supersonic combustion ramjet (scramjet) was search institutions around the world conducted many ground and flight
proposed by Antonio Ferri in the 1950s [3]. Based on this concept, the tests. The flight Mach number was increased and significant progress
engine configuration has been improved through research and the has been made.
structure was finally formed, as shown in Fig. 1. This engine mainly In supersonic combustion, there are interactions among a shock


Corresponding author.
E-mail address: wbing@tsinghua.edu.cn (B. Wang).

https://doi.org/10.1016/j.paerosci.2018.12.002
Received 28 June 2018; Received in revised form 4 December 2018; Accepted 18 December 2018
0376-0421/ © 2018 Published by Elsevier Ltd.

Please cite this article as: Ren, Z., Progress in Aerospace Sciences, https://doi.org/10.1016/j.paerosci.2018.12.002
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

wave, a turbulent vortex and a flame and strong coupling between


compressible turbulence and a chemical reaction, which are associated
with complex multiple temporal and spatial scales. The physical me-
chanism that completely explains these complicated phenomena is still
not well understood [6,7], and supersonic combustion in scramjet en-
gines is less well understood than low-speed phenomena. However, the
performance of supersonic combustion is found to largely depend on
the combustion characteristics of the fuel. Hydrogen and hydrocarbon
Fig. 1. Schematic diagram of scramjet engine structure [4]. fuels are commonly selected according to the flight Mach number and
the physicochemical properties of the fuel. Hydrogen fuel has a short
ignition delay time and a high heat release. The relevant research is
adequate. However, due to its small density under normal conditions,
the fuel tank for hydrogen should have a large volume in practical

Fig. 2. Schematic diagram of classic physical models employed for supersonic spray combustion: (a) mixing layer [8], (b) transverse jet [10], (c) backward-facing
step flow [11], and (d) cavity flow [12].

2
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

dispersed fuel sprays, droplet evaporation, fuel-air mixing and com-


bustion. Researchers have conducted many studies on the mixing layer
in which stability analysis, numerical simulation and experiments are
conducted. It is difficult to use linear stability analysis to explain the
nonlinear behavior of the mixing flow during its spatial-temporal evo-
lution. Experiments on supersonic flow are expensive and provide very
limited data. Numerical simulations have low cost and high efficiency
and the evolution of microscopic structures and the macroscopic sta-
tistical features of the flow field can be identified. However, high-fi-
Fig. 3. An illustration of the issues and interactions relevant to combustion of delity numerical solvers are essential for capturing flow and flame
liquid sprays [13]. structures in supersonic combustion. With the rapid development of
computational capabilities, numerical simulation has become an im-
applications. By contrast, liquid hydrocarbon fuel has the advantage of portant tool for studying supersonic flow and combustion.
easy storage, which is convenient for applications, and can also be used This paper reviews research efforts on studying the spray combus-
as the coolant. The combustion of liquid hydrocarbon fuel involves tion in supersonic flows in scramjet engines. First, background in-
complex physiochemical processes, such as atomization, evaporation, formation on supersonic spray combustion is introduced. Then, the
mixing of fuel and air and ignition. In addition, there are complex wave physical phenomena and characteristics that are involved in spray
structures and compressible turbulence in the combustion chamber of combustion are described and the related research results are sum-
the scramjet. The complexities of spray combustion and the flow marized. Aspects of the supersonic mixing layer, supersonic spray tur-
structures inside the supersonic combustor pose challenges for the de- bulent combustion, and shock-induced combustion are discussed with
sign and development of the scramjet engine. The investigation of su- an emphasis on ignition and combustion characteristics in supersonic
personic spray combustion focuses on atomization, dispersion of fuel flows.
droplets, evaporation, ignition and the interaction between the shock
wave and compressible turbulence. The identification of the corre- 2. Supersonic spray combustion
sponding mechanisms is of high academic significance and engineering
application value for developing the scramjet engine. The scenarios of spray combustion are complex, even if the atomi-
The flow and spray combustion in the supersonic combustor involve zation and droplet coalesce and collision are not considered. Jenny
interactions among turbulence, liquid spray, and a shock wave, along et al. provided a comprehensive review of the recent development in
with combustion. There are strong couplings between the multiple spray combustion that outlines the interactions (as shown in Fig. 3,
scales of these complex physiochemical processes, such as the multiple which is adopted from Ref. [13]) among the physical and chemical
scales in the turbulent vortex of the continuous phase, in the liquid processes. The research progress on the atomization [14,15], dispersion
sprays of the dispersed phase and in the spray combustion field. There of particles/droplets [16], ignition and flame [13,17] during spray
are also the interactions between the multi-physical and chemical combustion in subsonic flows was reviewed. For supersonic flows, the
processes, such as the couplings between the turbulence flow, heat and high-speed flow compressibility can result in strong flow compression
mass transfer, and combustion reaction. Therefore, it is necessary to and dilatation, which affects the combustion features of sprayed fuel.
clarify the mechanisms of the interphase couplings with multiple scales For instance, studies on the dispersion of particles in compressible
and to understand the interactions among the physical and chemical turbulent flows have demonstrated that the particles are transported by
processes in the inner/inter-phases. There are two main research ap- shocklets and preferentially accumulate in the high-vorticity regions
proaches to investigating the complex processes in supersonic com- behind the randomly formed shocklets [18–20]. Next, the research
bustion: One is based on the supersonic mixing layer, the supersonic jet progress on spray combustion in supersonic flows is reviewed.
and other basic physical models as the research objects. The corre- Due to the requirement of drag reduction in the internal passage,
sponding characteristics of supersonic combustion can be determined the combustion chamber in the scramjet must be very short [21].
via numerical simulation and experimental measurement. The other Hence, the associated residence time of the flow mixtures of fuel and air
focuses on the combustor model of the scramjet engine and the relevant is short in the combustor. Therefore, flame stabilization in supersonic
features can be determined via numerical simulation, ground experi- airflows becomes difficult to achieve and the rapid fuel-air mixing must
ments and flight testing. Due to the complexity of supersonic combus- be taken into account in engineering applications to ensure reasonable
tion, although many studies that are based on the two approaches have combustion efficiency. If the inflow temperature for the combustor
been carried out, the results of which are impressive, the physics of exceeds 1000 K, the auto-ignition becomes very fast. In this scenario,
supersonic combustion is still not well understood [6]. The first ap- the flame could be stable even if there is no flame holder, which is
proach is more fundamental and the associated conclusions might hold related to the physical and chemical properties of the fuel and its in-
universally. Not only can the basic phenomena and physical mechanism jection strategy. Struts, cavities and other fuel injection methods can
during supersonic combustion be identified via this approach but the lead to various recirculation zones, which affect the flame stabilization
results can also guide the engineering design and application of the and combustion efficiency. There are multiple time scales, such as the
combustor of the scramjet engine. mixing time of fuel and oxygen, the residence time and the chemical
The classical physical model of supersonic spray combustion in- reaction, during supersonic combustion. The associated time scale
cludes mixing layers, wall transverse/parallel jets, and backward-facing analysis is useful for estimating the characteristics of the combustion
step or cavity flows, as illustrated in Fig. 2. chamber. The supersonic flows inside the combustion chamber contain
As one of the classical physical models, the supersonic mixing layer both spanwise vortices and streamwise vortices (with a high degree of
is formed by the shearing of two coflows that have different speeds, as helicity). Although the barocline and expansion effects could attenuate
illustrated in Fig. 2 (a). The development of the mixing layer includes the turbulence kinetic energy, the presence of vortices enhances the
the basic characteristics of turbulence, such as flow instability, transi- mixing. In particular, the flow compressibility affects the turbulence
tion and the generation, pairing and merging of vortices; hence, it can intensity and the scaling law, in which the energy is transmitted from
be applied as a research model for studying the flow and combustion in large-scale to small-scale vortices, differs from that of the subsonic flow
the scramjet engine. In addition, the vortex dynamics in the mixing [21]. The turbulence kinetic energy of supersonic flows is found to
layer have significant effects on the atomization and motions of dissipate from lower wave numbers. Therefore, the size of the dissipa-
tion vortices in the supersonic flow becomes larger than that in the

3
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

subsonic flow, which has been demonstrated experimentally [22]. The equilibrium temperature [23], and the time that is required for the
flame thickness for supersonic combustion could be less than the temperature growth rate to attain its maximum value [25]. For in-
minimum scale of the vortices; if so, the minimum vortex would only stance, Mitani et al. [26] calculated that the ignition delay time for
wrinkle the flame and would not enter the flame and thicken it [21]. gaseous hydrogen fuel is in the order of 10−5 s.
Hence, in model selection for simulating the interactions between The burning time is defined as the time that is required to reach
chemical reactions and turbulence in supersonic flow, not only the 95% of the equilibrium temperature; this time is on the order of 10−3 s
Reynolds number but also the Mach number should be considered. The for hydrogen fuels [26]. Therefore, if molecular-level mixing has been
following part introduces the essential features of supersonic combus- realized, the reaction time and the residence time of the airflow in the
tion. Based on the classical physical model of the mixing layer, the combustor are of the same order. Accordingly, if there is no flame
current research status and development trend for supersonic flow and stabilizer to extend the residence time, it is not possible to complete the
combustion are discussed. exothermic reaction in the combustion chamber.
The main characteristics of supersonic combustion can be sum-
2.1. Basic features marized as follows: flow compressibility, strong couplings between
physicochemical processes, combustion instability, and turbulence-
As discussed above, supersonic combustion is complex and involves shock-combustion interaction.
physical and chemical processes with multiple scales, such as the
compressible turbulence, mixing between fuel and oxidant, chemical (1) The high Mach number of supersonic flows leads to strong flow
compressibility, which affects the development of large-scale vor-
reaction, and shock wave. The turbulence mixing is strongly coupled
with the chemical reaction due to their similar time scales. Combustion tices in the mixing process. With the increase of flow compressi-
bility, the mixing efficiency is found to be suppressed [27,28] and
initiates only when the mixing reaches the molecular scale; in turn, the
increase of temperature and the changes in the chemical components the baroclinic torque and dilatation become more important for the
entrainment process and the generation and development of flow
will further affect the local mixing.
There are a wide range of time scales during the physical and che- turbulence [29,30]. In addition, the flow compressibility also af-
fects the chemical reactions and the flame [31] due to the strong
mical processes in supersonic combustion. Assuming that the flight
compression and dilatation processes. Hence, Kelvin-Helmholtz (K-
Mach number of the aircraft is in the range of 6–8, the Mach number of
H) flow instability can result in the enhancement of mixing via the
the airflow in the combustion chamber is estimated to be between 2 and
generation of coherent flow structures. With the increase of flow
3 [23]. If the length of the combustor is on the order of meters, the
compressibility, the segregation of particles/droplets is found to be
residence time of the airflow is only on the order of milliseconds. Su-
transferred from the low-to high-vorticity regions and the shocklets
personic combustion must be completed efficiently during this short
in flows with high compressibility affect the preferential distribu-
time scale; hence, the supersonic airflow can generate heat to produce
tion of particles/droplets.
thrust. In particular, gaseous fuels must penetrate the air via injection
(2) Couplings between the physical quantities of pressure, density,
and further mix with the oxidant at the molecular level. The resulting
temperature and velocity occur in supersonic reacting flows. In
molecular collisions lead to the chemical reaction and hence the heat
release. For liquid fuels, atomization, evaporation and other complex particular, the increase of temperature is not only caused by the
heat release but is also related to the high-speed flow compressi-
processes must be considered. Fig. 4 shows the time scales for the re-
acting flow. Chemical reactions have a wide range of time scales, bility and viscous heating. Furthermore, the strong compression
and dilatation due to flow compressibility effects could result in a
namely, from 10−10 s to greater than 1 s. The time scales for molecular
transportation that is related to mixing range from 10−4 to 10−2 s. The significant change of pressure. The fluctuations of density and
species are caused by chemical reactions and strong fluctuations of
rapid chemical reaction process corresponds to the equilibrium state,
while the slow process corresponds to the frozen state. The chemical velocity and pressure in the supersonic flow.
(3) Auto-ignition of reactive mixtures is conducive to continuous and
reaction process can be decoupled from the flow field analysis under the
frozen state. However, this scenario does not exist during supersonic stable combustion. The high-speed airflow has a relatively high
enthalpy, thereby resulting in a relatively high static temperature in
combustion. From the perspective of chemical kinetics, the completion
of the exothermic reaction is related to two times: the ignition delay the local stagnation zone or low-speed region, which could con-
tribute to a rapid chemical reaction. At the same time, the strong
time and the time for completing combustion (the heat release time).
The ignition delay time has several definitions, such as the character- velocity fluctuations lead to uniform distributions of fuel gas and
temperature and flame extinction may occur during propagation in
istic time [24] that is deduced from the chain branching reaction rate,
the time that is required for the fuel-oxidant mixture to reach 5% of the the high-speed flow. Due to the interactions between the mixing
zone and the exothermic reaction zone, low-frequency oscillatory
combustion occurs in the combustion chamber and flow instability
can lead to high-frequency combustion instability [32].
(4) There are complex wave structures in the scramjet engine.
Expansion and compression/shock waves coexist. If a shock wave
interacts with the reaction zone or the flame, the temperature rises
due to shock compression and the temperature change that is
caused by the chemical reaction could be on the same order
[33–35]. Meanwhile, combustion will be intensified and the flame
structure will be affected by the shock wave. Shock waves can in-
duce the ignition of combustible fuel-oxidant mixtures, and a shock
wave may be coupled to the flame front if the conditions are ap-
propriate, which will have a positive effect on the flame stability. In
addition, the shock wave and combustion wave interact with tur-
bulent vortices, and the shocks contribute to increasing the vorticity
and mixing efficiency. The Richtmyer–Meshkov instability can re-
sult from the shock wave impacting the fluid/flame interface in the
Fig. 4. The time scales in the reacting flow [23]. scramjet combustor and can enhance the mixing process and

4
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

Fig. 5. Schematic diagram of mixing layer [9].

influence the ignition behavior.

For these characteristics of supersonic combustion, the present re-


view of supersonic spray combustion mainly considers the following:
(1) the atomization and mixing characteristics in supersonic flows; (2) Fig. 7. The effect of convective Mach number on the flow structure in the
the ignition, combustion characteristics and stabilization; and (3) the mixing layer [23].
interaction mechanism between a shock wave and combustion.
in the mixing layer. The K-H instability induces the vortex structure in
2.2. Supersonic spray combustion in mixing layers and jets the flow field. From Fig. 6(a)–(d), the values of the Reynolds numbers
increase and the value in Fig. 6 (d) is four times that in Fig. 6 (a). With
2.2.1. Supersonic mixing layers the increase of the Reynolds number, the small turbulence structures
The entrainment of vortices in mixing layers and jets could enhance become increasingly abundant. However, the large-scale vortices that
the mixing between fuel and oxidant in high-speed flows. In the mixing are induced by K-H instability in the flows that have high Reynolds
layer, the parallel flows have different speeds, as shown in Fig. 5, and numbers remain clearly visible.
the flows differ in terms of density and gas components. Momentum is For supersonic mixing layers, the flow compressibility has a sig-
transferred from the high-speed side to the low-speed side and mass is nificant effect on the flow dynamics. The density ratio of two streams,
delivered at the interface. Turbulent small vortices control the mixing velocity gradient, pressure gradient and heat release will affect the
and cause it to reach the molecular level; then, the chemical reaction development of the mixing layer. The convective Mach number, which
begins. For supersonic reacting mixing flows, it is necessary to in- is denoted as Mc, characterizes the flow compressibility and is ex-
vestigate the multiple scales that occur in the combustion process, pressed as,
which poses a challenge to experimental and theoretical research and
numerical simulation [36]. U1 − U2
Mc =
The flow structures in the mixing layer are of various sizes. The a1 + a2 (3)
Reynolds number, which is expressed as Re = δΔU/υ, affects the scale
of the vortex structure (where δ is the mixing layer thickness, ΔU is the where a is the speed of sound, and the subscripts 1 and 2 represent two
velocity difference between the two parallel streams, and υ is the ki- streams that differ in terms of velocity. Fig. 7 shows the effect of Mc on
nematic viscosity coefficient). Several researchers have proposed ex- the vortex scale in supersonic flows. The two-dimensional spanwise
pressions for the mixing-layer thickness growth rate. Slessor et al. [37] vortices in the flow with Mc = 0.525 are similar to the vortex structures
considered the compressible effects and expressed the growth rate as, in subsonic flows [23]. However, with the increase of Mc, the large-
scale vortices become more difficult to identify and the upper boundary
dδ (x ) 1 − ru ⎞ ⎛ 1 + rs ⎞ of the mixing layer becomes flat and exhibits almost no fluctuations.
= Cδ f1 (Mc ) ⎜⎛ ⎟ ⎜ ⎟

dx ⎝ 1 + rs ru ⎠ ⎝ 2 ⎠ When Mc increases to Mc = 1, the mixing area in the flow field becomes


(1 − rs )/(1 + rs ) ⎞ more uniform.
× ⎛1 −
⎜ ⎟
The increase of high-speed compressibility will reduce the turbu-
⎝ 1 + 1.29(1 + ru )/(1 − ru ) ⎠ (1)
lence intensity and suppress the growth rate of the mixing layer. These
where Cδ is the growth constant, which ranges from 0.25 to 0.45. The findings have been confirmed in previous experimental studies
empirical correlation, which is denoted as f1 (Mc) and represents the [40–42]. In particular, Samimy et al. [40] investigated supersonic
2
compressibility effect, is expressed as f1 (Mc ) = 0.8e−3Mc + 0.2 and mixing layers with Mc equals to 0.51, 0.64 and 0.86 using a laser
ru = UA2/UA1 and rs = ρA2/ρA1 are the free-stream velocity and density Doppler velocimeter (LDV). It was found that the turbulent intensity
ratios, respectively. Nuding proposed a more concise formula [38], in and growth rate of the mixing layer thickness were suppressed as the
which the growth rate is expressed in the following form, compressibility increased. The Reynolds stress is also related to the
(1 − ru )(1 + rs ) compressibility. In the initial stage of flow development, the large-scale
dδ (x )
= Cδ f2 (Mc ) vortices are intermittent. The high-order turbulent fluctuations are
dx 1 + rs ru (2)
suppressed with increasing Mc. Barre et al. [43] experimentally studied
2
where f2 (Mc ) = 0.7e−2.5Mc
+ 0.3 and Cδ = 0.14–0.17. a supersonic mixing layer with Mc = 0.62 and focused on the process
Fig. 6 shows the effects of the Reynolds number on the vortex scale from the turbulent boundary layer of the inlet wedge to the complete
development region. The effect of the turbulent Mach number, which is
denoted as Mt, was considered. It was demonstrated that the turbulence
diffusion in the shear flow is controlled by the large-scale structure, and
the effect of the compressibility on the velocity statistics (such as the
Reynolds stress) is stronger when Mc exceeds 0.6. Clemens et al. [44]
investigated the planar mixing layer with Mc in the range from 0.28 to
0.79 by utilizing planar laser Mie scattering (PLMS) and planar laser-
induced fluorescence (PLIF) based on concentrated alcohol clouds.
They found that the turbulence structure has two-dimensional features
Fig. 6. The effect of Reynolds number on the flow structure in subsonic mixing at low Mc and becomes three-dimensional when Mc exceeds 0.62. When
layer. Here, the Reynolds number increases from (a) to (d) [39]. Mc equals 0.28, the instantaneous distributions of the mixing fractions

5
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

have a gradient in the streamwise direction, but tend to be uniform in 2.2.2. Supersonic atomization
the spanwise direction. When Mc increases to the range from 0.62 to The atomization of a liquid jet that is subjected to the air streams is
0.79, the mixing fractions become uniform in the streamwise direction, an important factor that influences the combustion efficiency. The
but there is a gradient in the spanwise direction. These changes in the phenomenon of atomization can be briefly described as follows: When a
gradient from the streamwise direction to the spanwise direction de- liquid fuel jet is injected into a gaseous flow, K-H instability develops at
monstrate that the turbulent entrainments have different directions the gas-liquid interface, which induces R-T instability, thereby resulting
with different flow compressibility. The flow compressibility could re- in the formation of large droplets and ligaments. This process of dis-
duce the fluctuations of the mixing fractions (especially for the high- integration into large droplets and ligaments from the liquid jet is called
speed side of the mixing layers). Similar conclusions have been ob- primary breakup. The aerodynamic force from the relative inter-phase
tained from direct numerical simulation (DNS) studies. Vreman et al. velocities causes instabilities, thereby resulting in the further disin-
[45] considered the effects of compressibility (Mc = 0.2–1.2) on the tegration of droplets and ligaments into even smaller droplets; this
mixing layer via DNS. It was found that the growth rate decreases with process is called secondary breakup. In the supersonic flow, the ato-
the increase of Mc and the reduction of the pressure disturbance has a mization process of liquid fuel is complex and involves complicated
substantial influence on the change in the growth rate. Pantano et al. flow structures, such as bow shocks and shock-induced boundary layer
[46] applied DNS and found that the stress-strain terms in the Reynolds separation [50]. The secondary breakup of the liquid droplets also de-
stress transport equation decrease as the compressibility increases. This pends on the dimensionless parameters, such as the Weber number and
decrease not only hinders the transport of turbulence from streamwise the Ohnesorge number [51]. For a liquid fuel, the surface tension is
to spanwise directions but also suppresses the generation of turbulence, fixed; hence, the two-phase relative velocity determines the Weber
thereby reducing the turbulence intensity and the growth rate of the number, which mainly affects the secondary breakup of the droplets. As
mixing layer. The gradient Mach number, which is the ratio of the a result, in the supersonic crossflow, the two-phase relative velocity is
acoustic time to the fluid deformation time, determines the reduction of higher than in subsonic cases. Therefore, the relative velocity between
the pressure strain term. Moreover, if the density ratio of two streams is the two phases is much higher compared to the subsonic flow, which
larger than unity, the growth rate of the momentum thickness de- causes the secondary breakup play to a more important role. The small
creases. Atoufi et al. [47] investigated the kinetic energy exchange droplet size is the basis for the successful completion of combustion in
between the mean and fluctuating fields and analyzed the effect of supersonic airflow, which is also true for subsonic spray airflows.
compressibility on the production, pressure dilatation, and dissipation Knowledge of the droplet size distribution is vital for accurate spray
terms. The numerical simulations demonstrated that the production flame prediction.
term decreases linearly with the increase of Mc from 0.3 to 0.73, which, For liquid jets, four scenarios exist, which depend on the injection
in turn, leads to a linear reduction of the mixing layer growth rate. pattern of the nozzle: a round liquid jet that is discharging into quies-
Although the turbulent viscous dissipation term decreases with the in- cent gas, into a coaxial flow, and into a gas crossflow, and two im-
crease of the convective Mach number, it remains the primary kinetic pinging jets. For previous reviews on a round liquid jet that is dis-
energy dissipation mechanism at all convective Mach numbers. Wang charging into quiescent gas and into a coaxial flow, one can refer to
et al. [48] numerically studied the effects of the Mc, flow velocity ratio Bogy [52] and Lasheras & Hopfinger [53]. Recent studies on impinging
and fluid density ratio on passive scalar mixing. They found that the jets have been carried out by Chen et al. [54], Fu et al. [55], Rodrigues
mixing layer thickness and mixing efficiency both decrease as Mc in- et al. [56] and Zhang & Wang [57].
creases, as shown in Fig. 8. As the fluid density ratio increases, the In this review, we mainly focus on the atomization of a liquid jet in a
mixing layer thickness increases and the mixing efficiency of an en- supersonic gas crossflow. A schematic diagram of a liquid jet that is
trained fluid decreases. As the flow velocity ratio increases, the mixing discharging into a supersonic gas crossflow is shown in Fig. 9. As the
layer thickness decreases while the mixing efficiency increases. round liquid fuel jet discharges into the supersonic crossflow, a bow
The growth rate of the supersonic mixing layer decreases with the shock forms ahead of the jet. Meanwhile, the inverse pressure gradient
increase of the compressibility. When Mc increases, the reduction of in the streamwise direction induces a recirculation zone upstream of the
pressure fluctuations suppresses the interaction between the large-scale inlet of the jet and causes separation of the boundary layer. In the in-
vortices in the mixing layer and the two-dimensional instability mode is duced recirculation zone, the fuel and air mix and burn under subsonic
found to convert to an obliquely oriented instability mode [49], which conditions, which provides suitable conditions for flame stabilization.
is accompanied by decreases in the entrainment rates and the emer- The liquid fuel jet is atomized and forms small droplets; their higher
gence of three-dimensional features. Furthermore, the reduction of the penetration length will increase the mixing and combustion efficiency.
diffusion rate is related to the suppression of the production terms of Since the 1970s, many studies have been conducted experimentally
the turbulent kinetic energy. and numerically on the liquid jet breakup in supersonic gas crossflow.
Experimental approaches for studying the liquid jet in supersonic gas

Fig. 8. Mixing layer thickness and mixing efficiency of supersonic mixing layers for different values of the Mc [48].

6
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

Fig. 9. Schematic view of liquid jet discharging into supersonic gas crossflow
[50].

crossflow include photomicrographs, shadowgraphs, high-speed movies


[58], spark schlieren photographs [59], the pulsed shadowgraphic
technique [60], phase Doppler particle anemometry (PDPA) [61], di-
gital holographic microscopy [62], particle image velocimetry (PIV)
[63], and planar liquid laser-induced fluorescence [64]. In the experi-
ment, the main measured parameters are the spray penetration length
and the width of the spray plume. Scherman & Schetz [58] studied the
spray plume under gas inflow with Ms = 2.1 via photomicrographs and Fig. 10. Normalized centerline distribution profiles for SMD of pure-liquid jets
shadowgraphs. They observed that the small-amplitude unsteady wave in an M = 1.94 crossflow [65].
that develops quickly along the axial direction was the main mechanism
for the jet primary breakup. The injection diameter and dynamic and velocity in the near-injector region were obtained.
pressure and the liquid properties will influence the degree of breakup Numerical simulation is a complementary approach for the study of
and jet speed. Lin et al. [65] have conducted many experimental studies liquid jet breakup in supersonic gas crossflow. Liquid jet discharge into
on the spray structures of liquid jets. In their experiments, the pene- gas crossflow involves a complicated two-phase flow and the primary
tration height, cross-sectional area of the spray plume, flux-averaged and secondary breakup of the liquid jet vary substantially in both time
values across the spray plume and centerline distribution profiles were and space. Two-phase flow numerical methods such as VOF, Level Set,
considered. Based on their PDPA experimental studies, the following and Ghost Fluid are commonly used to study the primary breakup
correlation of the penetration height of pure-liquid jets was derived: mechanism of a liquid jet in the near field. Xiao et al. [69] used large
eddy simulation (LES) to study the primary breakup of liquid jets in
h 0 / d 0 = 4.73q00.30 (x / d 0)0.30 (4)
supersonic air crossflow. Fig. 11 shows the morphology of the liquid jet
where h 0 is the penetration height of the pure-liquid jets, d 0 is the in- and the pressure contours. The gas flow behind the liquid column,
jector orifice diameter, q0 is the jet-to-freestream momentum flux ratio, which is characterized by the effective Weber number, affects the sur-
and x is the distance in the freestream direction. They also derived the face waves on the windward side directly and the wavelength is well
correlation of the penetration height of an aerated-liquid jet as

(h − h 0)/ d 0 = 6.34q0−0.28 (x / d 0)0.30 (GLR)0.15 (5)

where h is the penetration of the aerated-liquid jet and GLR is the mass
ratio of the aerating gas to liquid, which is expressed as a percentage.
An “S” shape and a “mirrored-S” shape were observed for the normal-
ized distribution profiles, as shown in Fig. 10.
Recently, Regert et al. [66] used laser sheet imaging and infrared
light extinction spectroscopy (IR-LES) to study the breakup of liquid
ligaments in a Mach 6 crossflow. Their experimental results demon-
strated that the vicinity of the injection was influenced by the shapes of
the injector holes. Using the IR-LES technique, they found that droplets
of the smallest possible size were formed from each erupted liquid li-
gament since the catastrophic breakup type occurs. At this Mach
number, the mean droplet diameter is independent of the Weber
number. Lin et al. [67] have studied pure- and aerated-liquid jets in a
Mach 1.94 crossflow using the high-speed shadowgraph and PDPA
techniques. Their experimental results demonstrated that the initial li-
quid column breakup and eventual plume formation were substantially
influenced by the protrusion structure at the windward side of the in-
itial spray columns. The penetration height can be enhanced by a
convergent-divergent nozzle and an intrusion injector. For the aerated
jets, the droplet and plume properties of cross-sectional distributions
are highly similar. Samlam et al. [68] used inline digital holographic Fig. 11. Numerical results of the primary breakup of liquid jet (Ms = 2.1,
microscopy to study the near-field structures of pure- and aerated-liquid We = 594). The time interval between two adjacent subfigures is 2 × 10−6 s
jets in Mach 2 crossflows for the first time. The ligament/droplet size [69].

7
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

correlated to the effective Weber number. They found that the corre- structures are as shown in Fig. 12. The wave structures include the
lation between the Rayleigh-Taylor wavelength and the effective Weber incident shock wave (ISW), reflected shock wave (RSW), transmitted
number was the same as the correlation between the simulated surface wave (TW), reflected expansion wave (REW), re-transmitted wave (R-
wavelength and the effective Weber number; hence, the R-T instability TW), slip surface (SS) and Mach stem (MS).
plays an important role in the primary breakup of a liquid jet in a su- Recently, fully three-dimensional numerical simulations of shock
personic crossflow. The surface and column breakup are related to the interactions with droplets also provided insight into the breakup me-
vortices around the liquid jet. chanism of droplets. Liu et al. [88] performed numerical simulations of
In addition, researchers couple the Eulerian method and the droplet breakup under Ms = 1.2, 1.5 and 1.8. They concluded that the
Lagrangian method to study the primary and secondary breakup of the droplet breakup occurs under the shear-induced entrainment regime
liquid jet. Before the liquid jet breaks up into small-sized droplets, the based on their numerical results. The sources of the surface instabilities
two-phase interface flow is simulated via the interface tracking method. on windward side was the shear between the high-speed gas flow and
Then, as the liquid jet breaks up into small-sized droplets, the droplets the liquid droplet, while the instabilities on the leeward side were in-
are tracked via the Lagrangian approach [70–72]. When there are many duced by vortices. Meng and Colonius [89] performed a fully three-
droplets, it is difficult to track them in the Lagrangian framework. To dimensional simulation of droplet breakup under Ms = 1.47. The se-
overcome this drawback, the double-fluid method in the Eulerian fra- quence of deformations of the droplet are shown in Fig. 13. The nu-
mework was introduced to study the breakup process of a liquid jet in a merical results well agreed with the experimental results of Theofanous
supersonic gas crossflow [73]. [74]. They proposed that the droplet breaks up in the stripping breakup
regime. The observed surface instabilities provided qualitative evidence
2.2.3. Interaction of shock waves and droplets for the increase in the R-T instability along the accelerated sheet.
Recently, Theofanous [74] comprehensively reviewed the aero-
breakup of droplets; readers can find more details on the breakup 2.2.4. Supersonic spray combustion in mixing layers and jets
mechanism of droplets in this review. In this part, we will focus on the Due to technical reasons and cost constraints, the experimental
interaction of shock waves with droplets since there are strong inter- studies of supersonic combustion are insufficient. Therefore, the ex-
actions between shock waves and droplets during the secondary perimental data are scarce and the effect mechanism of ignition and
breakup of liquid jets in a supersonic crossflow. heat release on the mixing layer development is unclear. Heat release
Regarding the interaction of a shock wave with a droplet, the dro- can cause a decrease in density and an increase in volume. For a rela-
plet deformation and breakup mechanisms and shockwave interference tively low-speed flow (Mc < 0.4), Hermanson et al. [90] found that the
have been investigated for decades. The classical theory that describes mixing-layer growth rate decreases slightly as the heat release in-
droplet breakup was developed by Hanson et al. [75], Pilch & Erdman creases, which is related to the decrease of the turbulent shear stress.
[76], Joseph et al. [77], and Hsiang & Faeth [78–81]. Under low-speed The average distance between the large-scale coherent vortices de-
(subsonic) flows, the breakup mechanism can be classically summar- creases with the increase of temperature and vortex merging is in-
ized in five regimes according to the Weber number We, namely, vi- hibited by the chemical heat release, as shown in Fig. 14.
brational breakup, bag breakup, bag-and-stamen breakup, sheet strip- For a high-speed compressible flow, the effect of heat release is
ping and wave crest stripping, followed by catastrophic breakup. weaker [91]. Miller et al. [92] studied the effect of compressibility on
Due to the inherent limitation of the experimental approach, it is the hydrogen-air reacting mixing layer via OH-acetone plane laser-in-
difficult to identify detailed flow structures for shock interactions with duced fluorescence (PLIF) for Mc values of 0.32, 0.35 and 0.7. Under
liquid droplets. Instead, researchers have conducted experiments on moderate heat release, the vortex structure in the reacting flow is si-
shock interactions with liquid cylinders to identify the flow character- milar to that in the non-reacting flow for a similar compressibility and
istics. Igra and Takayama [82] used double-exposure holographic in- the mixing layer growth rate does not change substantially. In addition,
terferometry to qualitatively visualize the gas and liquid phases. Wave they have shown that under their research conditions, the compressi-
interactions and density variations were examined in their experiments. bility has a positive effect on the combustion in the mixing layer and the
They also conducted numerical simulations for the same cases as in the enhancement of the combustion is due to the change of the entrainment
experiments and the numerical simulation results well agreed with the process from incompressible to compressible flows. The numerical in-
experimental findings [83]. Recently, Sembian et al. [84] conducted vestigations by Calhoon et al. [93] confirmed that the chemical heat
shock interaction experiments with a cylindrical water column via the release decreases the growth rate and suppresses the mixing features of
high-resolution shadowgraph technique. The diameter of the cylinder the mixing layer. Mahler et al. [94] found that the compressibility and
was 22 mm and Mach numbers of 1.75 and 2.4 were examined. From heat release had the same effect on the flow of the mixing layer in their
their experimental results, various detailed flow features were identi- DNS studies, which demonstrated that the increase of Mc and the che-
fied, such as focusing of expansion waves, nucleation of cavitation mical reaction can decrease the growth rate and reduce the Reynolds
bubbles, and a re-circulation zone, that had not been identified in stress and the pressure-strain correlation. The decrease of the turbulent
previous experiments [85] and numerical simulations [86,88,89]. kinetic energy and Reynolds stress due to heat release are related to the
Xiang and Wang [87] performed numerical simulations of shock in- significant reduction of the gas density around the flame front. Mathew
teractions with the cylindrical water column via a highly accurate nu- et al. [95] studied the effects of compressibility and heat release on the
merical scheme and obtained the same results as in the experiments of entrainment process in the mixing layer via DNS and found that both
Sembian et al. [84]. According to the experimental results of Sembian compressibility and heat release impeded the entrainment process, as
et al. [84] and the numerical results of Xiang and Wang [87], the wave shown in Fig. 15.
Ferrer et al. [96] conducted a DNS study on the three-dimensional
spatially developing high-speed H2/air mixing layers and demonstrated
that the growth rate decreased according to the pressure-dilatation.
They found that the decay of pressure fluctuations may not be the sole
reason for the reduction of the growth rate, which is also related to the
decrease of strain-rate fluctuations. The influence of engulfment, which
plays an important role for flows with low compressibility, tends to
disappear in the presence of heat release and under increasing com-
Fig. 12. Schematic diagram of the wave structures at the early stage of a planar pressibility. The effect of heat release on the supersonic flow is same as
shock wave interacting with a cylindrical water column [87]. that of the flow compressibility, which could inhibit the turbulence

8
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

Fig. 13. The aero-breakup of a water droplet [89].

Fig. 14. Schlieren photographs at low (up) and high (down) heat release of
supersonic mixing layers [90].

intensity and the mixing layer development [94], and is primarily due
to the sharp decrease in the mean density around the flame sheet;
therefore, heat release is mainly a mean density effect. Researchers
from the Spray Combustion and Propulsion Laboratory in Tsinghua
Fig. 15. Iso-surfaces of oxygen concentration from nonreacting cases (a)–(c)
University thoroughly investigated the supersonic combustion in
and mixture fraction from reacting cases (d)–(f). (a) Mc = 0.15, (b) Mc = 0.7,
mixing layers. They constructed a single-step H2/O2 chemical reaction (c) Mc = 1.1. (d) Mc = 0.15, reacting, (e) Mc = 0.7, reacting, (f) Mc = 1.1, re-
mechanism for predicting ignition in supersonic flows [97]. Zhang et al. acting [95].
[98] found the unsteady phenomena of ignition, flame propagation and
extinction in the supersonic mixing layer. Chen et al. [99] analyzed the
2.3. Stabilization approaches
H2/air supersonic combustion mode in the mixing layer and found that
competition between vortex shedding and auto-ignition led to com-
Compared with hydrogen fuels, liquid hydrocarbon fuels have
bustion instability. Ren et al. [100] investigated spray combustion in a
higher density and are easier to store; hence, the fuel tank has a rela-
droplet-laden supersonic mixing layer. They demonstrated that an ig-
tively smaller volume. At the same time, hydrocarbon fuels are safe and
nition kernel occurs in the high-strain regions and extinction resulted
of low cost; hence, they have a relatively wide range of applications. In
from the vortex dynamics and preferential dispersion of droplets. The
addition, liquid hydrocarbon fuels are typically used as endothermic
evolutions of flame kernels over one eddy turnover time are illustrated
coolants and their associated combustion characteristics, such as igni-
in Fig. 16. In addition, Ren et al. [101] found that there was an optimal
tion delays and flame propagation velocities, are improved. The main
size for droplets transported by the shear vortex for which the ignition
challenge for spray combustion in the scramjet combustor is the com-
delay time was the shortest and the flame temperature attains its
pletion of atomization, evaporation, mixing, ignition and combustion
highest value, as shown in Fig. 17.
within a short time. Compared with hydrogen fuel, the atomization and
evaporation of liquid hydrocarbon fuel must be considered, and the

9
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

The deep penetration of atomized liquid fuel into the supersonic


airflow is conducive to mixing, and important for maintaining the
combustion. For a combustor of the scramjet engine with a flight Mach
number from 6.0 to 7.0, the typical penetration depth of a fuel jet is
approximately 10–15 mm. Experimental and numerical studies have
demonstrated that the chemical reaction zone that results from the fuel
injection of the combustion chamber wall occupies only a small portion
of the complete flow field. Thus, not all the oxygen that is supplied by
the supersonic flow can participate in the chemical reaction process and
if the reaction zone is near the wall, it will impose additional thermal
loads on the engine structure. For flame stabilization systems such as
wall injection, step, ramp, cavity, and strut, could produce a low-speed
flow separation zone or vortex to stabilize the chemical reaction zone,
thereby improving the combustion performance, as shown in Fig. 18.
Fig. 16. Evolution of flame kernels within one eddy turnover time [100]. The injection of fuel into the vortex through a strut or another in-
jector could extend the residence time in the combustion chamber.
Furthermore, the oblique shocks that are generated by the interaction
atomization is particularly important, as it determines the spatial dis-
of the supersonic inflow and the strut enhance the mixing of fuel and
tribution of the fuel droplets and affects the subsequent combustion.
air. Fuel injection based on the strut has been experimentally studied in
The ignition delay time of the fuel is longer than the residence time (in
a small-sized model combustor and tested under flight Mach numbers of
milliseconds) of the airflow in the combustion chamber. In addition, the
4, 6 and 8 [104]. Many experimental and numerical investigations have
liquid fuel must be atomized, evaporated to vapors, and mixed with the
focused on the resistance loss, mixing and combustion in the combus-
oxidant for the chemical reaction, which could require additional time.
tion chamber of a scramjet engine for strut fuel injection. Manna et al.
Therefore, it is typically necessary to construct a recirculation zone near
[105] studied the combustion characteristics of the liquid fuel that is
the igniter or the fuel injector to prolong the residence time of the
injected by the strut in the supersonic flow and considered the influence
airflow, thereby ensuring the completion of ignition and combustion.
of the inflow Mach number and the total pressure at the inlet of the
The recirculation zone also acts as a flame stabilizer; however, it in-
combustion chamber. It was found that the drag that is induced by the
troduces additional total pressure loss. The design of the flame stabi-
strut has important effects on the performance of the combustor and the
lizer not only reasonably extends the residence time of the airflow but
setting patterns of the injection holes affect the induction process of the
also ensures that the pressure loss is in the affordable range. The
streamwise vortices. Kumaran et al. [106] extended Manna's research
combustion stability depends on the stable combustion limit of the fuel
and found that segmented injection could enable more fuel to partici-
[102], which is related to the temperature, velocity, pressure and
pate in the combustion without changing the inlet conditions. Vino-
equivalent ratio. The stable combustion limit of hydrogen fuel is larger
gradov et al. [107] experimentally studied the ignition and flame sta-
than that of hydrocarbon fuel. Among the hydrocarbon fuels, methane
bility of kerosene fuel, which is injected from the strut, and found that
has the narrowest range and cracking could broaden the range of the
stable combustion could be realized, even without the assistance of
stable combustion limit.
hydrogen fuel. Kouchi et al. [108] numerically investigated the

Fig. 17. Instantaneous distributions of the dimensionless gaseous temperature for four cases at a same time, and the initial droplet size increases from (a) to (d)
[101].

10
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

Fig. 18. Typical flame stabilizers in combustors [103].

Fig. 19. Schematic of flowfield structure for scramjet mode, weak ramjet and strong ramjet mode [109].

transition of the combustion mode in a supersonic combustor based on the double-strut configuration was applied. Yang et al. [110] found that
strut injection and found that four combustion modes, namely, extinc- the first-stage strut has a larger effect in the subsonic combustion mode,
tion, weak combustion, strong combustion and thermal choke, occurred while the second-stage strut plays a more important role in the super-
with the variation of the fuel flow rate. In the weak combustion mode, sonic mode.
most of the reactive mixtures are not ignited and the combustion zone is Compared with other flame stabilization systems, the wall cavities
confined near the wall of the chamber. In the strong combustion mode, have less influence on the loss of drag and the total pressure. Yu et al.
the diffusion flames appear and are anchored under the influence of the [111] studied the fuel injection and flame stability features in the su-
rear-step structure. The combustion mode transition was found to be a personic model combustor using liquid kerosene and considered the
direct function of the ignition capabilities, with the weak combustion effects of the fuel injection angle and the cavity configuration on the
confined to the downstream boundary layer. Zhang et al. [109] de- flame stability and the mixing enhancement. It was demonstrated that
monstrated that the combustion mode in the scramjet engine could be the addition of hydrogen significantly improved the combustion effi-
divided into the scramjet mode, the weak ramjet mode and the strong ciency and the combination of open and closed cavities yielded better
ramjet mode according to the equivalence ratio, as shown in Fig. 19. In combustion performance than a single cavity. By adjusting the fuel
their experiments, they identified the mode transition by monitoring injection position, the amount of hydrogen addition can be varied ac-
the wall pressure and found that the transition has a strong non-line- cordingly. In addition, Yu et al. [112] investigated the combustion
arity, in which the static pressure exhibits a discontinuous, sudden characteristics and flame structure of the cavity via planar laser-in-
change as the transition occurs. To strengthen the combustion process, duced fluorescence (PLIF) and schlieren photography. They observed

11
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

inconsistencies in the characteristics of the cavity between non-reacting stabilization zone via strut injection to improve the combustion effi-
and reacting supersonic flows. Kumaran et al. [113] numerically si- ciency. In general, it is appropriate to improve the supersonic com-
mulated the mixing and combustion process of kerosene in a model bustion efficiency and realize the stable combustion by applying a
combustor via Reynolds-averaged Navier-Stokes (RANS) simulations multi-cavity configuration in combination with a reasonable fuel in-
that were coupled with a single-step chemical reaction model. It was jection method. However, the strategy of combing cavities with other
found that the penetration depth, the dispersion of fuel and the eva- injectors and the transition of combustion modes require further, in-
poration rate increased with the increase of the injection pressure, depth study.
which was associated with increases in the mixing efficiency and the Numerical investigations of the flow and combustion processes in
combustion efficiency. The turbulent dispersion of fuel droplets had a the combustor of the scramjet engine mainly consider hydrogen or
significant effect on the heat release. Pan et al. [114] experimentally gaseous hydrocarbon fuels with small molecules. For supersonic com-
studied the supersonic combustion characteristics of kerosene for bustion using liquid hydrocarbon fuels, it is necessary to consider not
double cavities and found that it was easier to stabilize the flame when only the atomization and evaporation processes but also the supersonic
the cavities are in the serial configuration. Tsue et al. [115] investigated turbulent two-phase reacting flows with multiple scales that are asso-
the characteristics of auto-ignition and flame stability for various hy- ciated with the complex chemical reaction mechanisms for hydro-
drocarbon fuels (the number of carbon atoms ranges from 7 to 16) in carbon fuel, which are challenging to simulate numerically.
supersonic flows. It was found that the fuel with the fewest carbon
atoms yielded the best auto-ignition performance and the volatility of
2.4. Shock wave and combustion
the fuel had a crucial influence. The fuels that had 8 to 10 carbon atoms
yielded the highest flame stability. Bao et al. [116] studied the com-
2.4.1. Effects of shock wave on combustion
bustion characteristics of liquid kerosene in double cavities via spark
To enhance the mixing of fuel and air to improve the combustion
ignition and the influence of the cavity length on the ignition was
efficiency, mixing flows are typically used in a supersonic combustor.
analyzed by comparing the wall pressure and thrust. They found that a
As the inflow Mach number increases, it becomes difficult for the
long cavity (length-to-depth ratio L/D = 7) was more suitable for spark
mixing layer to lose its flow stability and the mixing efficiency de-
ignition than a short cavity (L/D = 5). For the short cavity, there were
creases. Therefore, mixing enhancement has been attracting the atten-
three stable combustion stages, in which the equivalence ratio for the
tion of researchers. Both active and passive approaches have been
shear layer of the cavity determines the evolution of the local flame and
proposed for strengthening the mixing of supersonic flows in recent
the final stable combustion depended on the flame development pro-
decades [123]. One of the passive methods, namely, the introduction of
cess. Gang et al. [117] applied RANS to compare the effect of the re-
oblique shock waves into the flow field, has attracted more attention
action mechanism on the supersonic combustion in triple cavities. Ac-
from researchers because shock waves are inherent flow structures that
cording to the experimental results, the single step reaction could
are generated in the combustor of scramjet engines. Researchers expect
capture the changes in the wall pressure of the combustor; however, the
the utilization of shock waves to become an economical and effective
multi-step reaction model provided more details. Wang et al. [118]
approach to promoting supersonic mixing. Kumar et al. [124] produced
determined the combustion efficiency of a combustion chamber with
oscillatory shock waves for mixing augmentation in scramjet combus-
double cavities via tunable diode laser absorption spectroscopy
tors. Marble [125] proposed a method for enhancing the mixing rate
(TDLAS) and found that increasing the expansion angles could reduce
and combustion via the introduction of streamwise vorticity that is
the thermal choke and that the high flow velocity led to incomplete
generated by the interaction of an oblique shock with the density gra-
chemical reactions. In addition, the double cavities must be combined
dient between air and hydrogen, as depicted in Fig. 20.
with a multi-injector for operation, in which the downstream injector
Huh H et al. [126–128] found that shock waves enhanced the
with cavities could increase the penetration depth of the fuel and
mixing of fuel and air and improved the combustion stability limit since
strengthen the mixing. An arrangement that leaves a reasonable dis-
a shock wave can produce vorticity and an adverse pressure gradient,
tance between the cavities could stabilize the flame and the application
which increased the entrainment rate and mixing rate at the root of the
of double cavities enlarges the recirculation zone, leading to satisfac-
flame. Li et al. [129] found that an oblique shock wave could induce
tory mixing. Zong et al. [119] experimentally studied the supersonic
significant small-scale turbulence and that the intensity of the shock-
combustion of kerosene based on the combination of a cavity and a
generated turbulence increased with the oblique shock angle, α, as
strut and found that the increased amount of fuel that was injected from
shown in Fig. 21.
the strut promoted the thrust performance. Bao et al. [120] identified
Kim et al. [130] found that the growth rate of the shear-layer
the effects of partially covered cavities on the spark ignition of kerosene
thickness substantially increased after an interaction with an oblique
and analyzed the effects of the injection pressure and length of the
shock wave; hence, the mixing efficiency was found to increase in su-
cover plate on the ignition and flame propagation. Nakaya et al. [121]
personic airflows. Génin and Menon [131] studied a mixing layer that
experimentally studied the combustion characteristics of ethanol fuel in
a supersonic cavity. They measured the droplet diameter distribution by
using a laser diffraction spray analyzer (LDSA) to estimate the eva-
poration times of droplets and photographed the CH radicals and
schlieren graphs, from which they identified two combustion modes: a
strong combustion mode and a transition mode. The combustion per-
formance of ethanol was between those of the ethane and ethylene. In
addition, they estimated the ignition limit by calculating the liquid
breakup, evaporation, residence time and chemical reaction time. Al-
though the increase of the penetration depth promoted fuel-air mixing,
a penetration depth that exceeded 10 mm could result in the absence of
supersonic combustion, with an equivalence ratio that was equal to
unity. Sun et al. [122] used multiple cavities to study the stable com-
bustion of supercritical kerosene in supersonic combustors. The fuel-air
mixing could be enhanced via intrusive injection by thin struts; how-
ever, the combustion efficiency was not significantly improved. They Fig. 20. Progression of shock through the hydrogen jet to produce streamwise
concluded that it was feasible to establish a continuous flame vortex [125].

12
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

increases in the mixing layer thickness and the mixing efficiency.


Lee et al. [133] studied the shock-enhanced mixing and combustion
of hydrogen in supersonic flows via numerical simulations (RANS
coupled with detailed reaction mechanisms) and focused on the inter-
actions of oblique shocks and streamwise vortices that were induced by
ramps. They found that the chemical heat release had a slight effect on
the generation of vorticity, whereas the oblique shock wave and the
associated density gradients determined the vorticity. Mixing strongly
influenced combustion; however, combustion did not substantially af-
fect mixing. Roy et al. [134] numerically investigated the effects of an
oblique shock that was induced by a wedge on flame stabilization and
found that the action of the shock on the recirculation zone induced
Fig. 21. The instantaneous pressure contours in the supersonic mixing layer intense mixing, thereby resulting in a longer residence time. The as-
interacting with an incident shock wave with different shock angles [129]. sociated high density and temperature promoted flame stabilization.
Ratner et al. [135] experimentally studied the effects of oblique shocks
on the combustion efficiency of supersonic jet flames. Due to the
was affected by two oblique shock waves and found that the turbulent
changes in the local flow structures that are affected by shock waves,
fluctuating levels enlarged and the growth rate increased as the inter-
the flame becomes narrower and local extinction occurs, thereby re-
action position moved downstream; however, they also observed that
ducing the combustion efficiency. Nakamura et al. [136] investigated
the mixing enhancement is localized around the action position of the
the interactions of shock waves and transverse jets of hydrogen via
shock wave on the mixing layer. Zhang et al. [132] found that the in-
numerical simulations and experiments and analyzed the effects of the
troduction of an oblique shock resulted in the enhancement of the
incident position of the hock. They demonstrated that the shock wave
fluctuating levels of velocities and turbulent kinetic energy, leading to
widened the separation zone upstream of the fuel injection hole and

Fig. 22. Schematic view of different post-shock configurations [140].

13
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

reduces the penetration depth. When the total temperature of the in-
coming flow becomes lower, a flame stabilization zone can form under
the action of the shock wave. Mai et al. [137] experimentally studied a
non-reactive flow via planar laser-induced fluorescence of nitric oxide
(NO-PLIF) and particle image velocimetry (PIV) and investigated the
effect of an incident shock on the flow structure and mixing via three-
dimensional simulation of reactive flows. It was found that flame sta-
bilization could be achieved only when the incident shock was in-
troduced in a downstream regime of the injector. Shekarian et al. [138]
numerically considered the effect of incident shock waves on the flame
stability of transverse jets. They studied the effect of the incident po-
sition of the shock on the mixing and flame stabilization. As the in-
cident shock moves downstream from the fuel injector, the size of the
recirculation zone and the mixing efficiency increase significantly,
thereby promoting flame stabilization. Hariharan et al. [139] numeri- Fig. 23. Schematic diagram for the normal shock induced combustion [144].
cally analyzed the effects of the total temperature and the Mach number
on supersonic combustion. It was found that when the oblique shock
acted on the downstream regime of the transverse jet, the size of the
recirculation zone increased, whereas the total temperature had a weak
effect on combustion. As the inflow Mach number increases, the in-
crease of the fuel mass flow rate entering the mainstream promotes
fuel-air mixing. Huete et al. [140] studied the ignition in a supersonic
laminar mixing layer with an incident oblique shock by solving the
linearized Euler equation. They found that an ignition kernel occurred
in the highest-temperature regime, which could be located at the edge Fig. 24. Schematic diagram for the oblique shock induced combustion [145].
or the interior of the mixing layer, depending on the distribution of the
fuel concentration, the temperature and Mach number, and that the
normal shock and the pressure and temperature increases across the
formation of an ignition kernel resulted from the competition between
oblique shock can be adjusted by changing the oblique shock angle. The
the chemical reaction and shock-induced fluid expansion, as illustrated
application of oblique shocks for inducing ignition is very attractive.
in Fig. 22. They also applied a two-step reaction of hydrogen to in-
Rhodes and Rubins [145] injected hydrogen in the upstream regime of
vestigate the ignition of diffusion flames [141], and found that the
an oblique shock, as shown in Fig. 24. It was found that the hydrogen
chemical heat release and the acoustic effects from shock waves make
and oxygen content in the downstream regime of the oblique shock
the shock surface more curved.
gradually decreased; hence, combustion occurred.
Ren et al. [33] analyzed the interactions of a large-scale vortex, an
In supersonic flows, if the oblique shock wave is sufficiently strong,
incident oblique shock wave, and a chemical reaction in a supersonic
the chemical reaction zone that is formed in the fuel-air mixture that is
shear layer that was laden with droplets and found that two reaction
passing through the shock wave could be coupled with the leading
modes could be distinguished during the shock-vortex interaction: a
shock, thereby resulting in the occurrence of a detonation wave.
thermal mode and a local quasi-detonation mode. Studies of shock-in-
Kasahara et al. [146] generated a steady-state detonation around a
duced combustion in supersonic flows mainly focus on the overall
features, such as the recirculation zone and the combustion efficiency.

2.4.2. Shock-induced combustion


The complexity of gas dynamics and chemical reaction kinetics in
supersonic combustors poses problems for the design of supersonic
propulsion systems [142]. The shock waves that occur in the combus-
tion chamber could accelerate the mixing of chemical components
[143] and promote combustion [142]. From the perspective of che-
mical reactions, the main effect of a shock wave is a sudden compres-
sion, which increases the post-shock gaseous temperature. If this tem-
perature exceeds the fuel ignition temperature, the chemical reaction of
fuel-air mixtures will initiate quickly.
As early as the 1960s, Nicholls et al. [144] began to induce com-
bustion via the generation of normal shocks. Their experimental
equipment is shown in Fig. 23. Cold hydrogen was injected into su-
personic hot air. When the reactive mixtures passed through the normal
shock, various phenomena were observed, such as supersonic com-
bustion and standing detonation waves. As the airflow passes the
normal shock, a high inflow Mach number will result in high tem-
perature and pressure of post-shock flows, which will lead to high
thermal loads, extreme cooling problems, and chemical cracking. These
effects could delay the completion of the exothermic reaction and, thus,
reduce the thrust. Therefore, combustion that is induced by a normal
shock is not recommended. Oblique shock waves can also promote
combustion. It is necessary to have a higher total temperature for the
airflow to produce the required static temperature for ignition since the Fig. 25. Oblique detonation wave around hypersonic projectile with projectile
temperature passing through the oblique shock is less than that of the velocity varied from 2.9 to 4.1 km/s [146].

14
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

experimentally studied the detonation waves that were generated by a


hypersonic spherical projectile in an explosive mixture [148] and found
that the wave structures consisted of a shock-induced combustion (SIC),
a stable oblique detonation wave (ODW), and a wave of straw-hat type,
which consisted of a strong SIC and ODW. They [149] found that the
stabilizing criticalities of ODW depended on the ratio of the projectile
diameter and the cell size of the mixture. In addition, they [150] pro-
duced high-time-resolution schlieren visualizations to show whole
processes, from unsteady initiations to stable propagations of the sta-
bilized ODW, as depicted in Fig. 27. The effects of the projectile scale on
the stabilization of ODW were also considered [151], and the projectile
radius (diameter) was found to proportionally affect the geometrical
scale of the wave around the projectile.
Iwata et al. [152] numerically studied the effects of the nonuniform
premixing of a hydrogen-air mixture on the wave structures around a
hypersonic spherical projectile and found that the increase of the
nonuniformity changed the determinant factors in the flame structures.
Researchers have often generated an oblique shock by passing a
supersonic premixed fuel-air flow over a wedge. The oblique shock
compresses the mixture gas to induce chemical reaction and detonation.
Then, the concept of the oblique detonation wave engine (ODWE) was
then proposed [153]. Pratt et al. [154] studied the effects of the wedge
angle on the initiation of detonation waves based on the shock polar
analysis. It was found that the uncompleted oblique detonation, ob-
lique-shock-induced combustion and non-combustion modes occurred
when the wedge angle is small and the oblique shock is separated from
the wedge. Overdriven normal detonation or normal-shock-induced
combustion occurred when the wedge angle became large. Li et al.
[155] examined, in detail, the structure of oblique detonation via nu-
merical simulation and found that the detonation consisted of four
parts, namely, a non-reacting shock, an induction zone, a series of de-
flagration waves and a reacting shock (the shock front and the energy
Fig. 26. Oblique detonation wave around hypersonic projectile with initial
release zone are coupled), as shown in Fig. 28. They found that the
pressure varied from 41.3 to 66.3 kPa [147].
detonation wave structure was stable over a wide range of flow and
mixing conditions.
hypersonic projectile in stoichiometric hydrogen-oxygen premixed Thaker et al. [156] studied the effect of the reaction mechanism and
gases and visualized the flow field around the projectile, as shown in demonstrated that the two-step mechanism was suitable for the time
Fig. 25. They also illustrated the detailed structures for stabilized ob- scales in the physicochemical process of oblique detonation. Choi et al.
lique detonation waves around spherical bodies in a mixture of acet- [157] considered the effects of the fuel activation energy on the shock-
ylene, oxygen and krypton [147], as shown in Fig. 26. Maeda et al. induced detonation wave structure via numerical simulation and found

Fig. 27. Continuous pictures of the initiation process of stabilized ODW [150].

15
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

simulations, and the effects of inflow parameters and chemical re-


activity of fuel on the shock-to-detonation transition and the detonation
wave structure have been considered. However, almost all the present
studies were based on the gaseous fuel and a deep knowledge for the
shock-induced oblique detonation using liquid hydrocarbon fuel is still
lacking and must be fully addressed in the future.

Fig. 28. Schematic of the basic structure of oblique detonations generated by 3. Challenges and prospective
wedge-induced shocks [155].
According to the above discussion of various phenomena, numerical
that the detonation wave changed from a stable front to a cell-like wave simulation approaches and experimental databases, previous in-
front as the activation energy of the reaction increased. Verreault et al. vestigations have provided significant insight into supersonic spray
[158] found that a shock-induced Chapman-Jouguet (C-J) detonation combustion, which is a very complex physiochemical process. The large
could occur even if the wedge angle is less than the required deflection amount of research performed in recent years has improved the un-
angle for an oblique C-J detonation. Gui et al. [159] numerically stu- derstanding of physics of supersonic spray combustion. However, su-
died an oblique detonation wave for a Mach 7 inlet flow over a suffi- personic multiphase combustion physics is still not well understood.
ciently long wedge and identified three regions in the flow field behind This review highlights only a small portion of these breakthroughs.
ODW: a Zeldovich-von Neumann-Doering (ZND) model-like structure, a There are other relevant aspects of supersonic spray combustion in
single-sided structure with a triple point, and a dual-headed structure scramjets that are not addressed here, such as plasma-assisted ignition
with a triple point. The transition from oblique shock to detonation methods for supersonic flows [172] and highly resolved spatiotemporal
through a triple wave point (abrupt transition) was demonstrated both data that are obtained using advanced experimental diagnostics [173],
experimentally and numerically [160] and was considered a standard which, in conjunction with theoretical analyses and numerical simu-
structure [161]. In addition to this abrupt transition structure, which lations [174], can lead to substantial developments in this research
has a triple wave point, a smooth shock-to-detonation transition from a field.
curved shock was observed. Teng et al. [162–165] numerically con- In most practical scramjet configurations and operating conditions,
sidered the transition pattern of the oblique detonation and identified flame stabilization typically does not solely rely on autoignition, par-
abrupt and smooth transition patterns, as shown in Fig. 29. The inflow ticularly for liquid hydrocarbon fuels that have long ignition times.
conditions influence the transition structure and the shock-detonation Ignition and flame stabilization are realized via shock-induced me-
transition modes depend on the inflow Mach number, activation energy chanisms and the recirculation zone under appropriate injection con-
and heat release. They also studied the effects of the inflow pressure figurations such as struts and cavities. Once ignited, diffusion flames
and Mach number on the structure and length of the induction zone tend to persist in the combustor, even under supersonic coflowing
[166]. It was found that the induction length increased with the in- speeds, and shock impingements and flow compressibility effects may
crement of the inflow Mach number, whereas the effect of the inflow alter the mixing dynamics and flame structures. For atomized fuel
pressure was slight. droplets with multiple scales, these effects are able to change the flow
Zhang et al. [167] investigated the influence of the equivalence patterns due to turbulence modulation effects, which also pose chal-
ratio and the inflow speed on the induction zone that connects the lenges in the study of supersonic turbulent combustion.
oblique shock and the detonation waves and found that the induction Several aspects are worthy of future investigations; these challenges
length was shortest when the equivalence ratio was 0.8 for hydrogen are listed as follows:
fuel. The inhomogeneous distribution of reactive mixtures could also
affect the wave structures [168] and V-shaped deflagration and V + Y (1) A fundamental understanding of the ignition and flame stabilization
Mach stem could occur, as illustrated in Fig. 30. in supersonic reacting flows with fuel sprays must be obtained,
The effects of mixture equivalence ratio inhomogeneity were also especially via shock-induced mechanisms because their associated
studied by Fang et al. [169]. They found that the initiation length was autoignition times are considerably longer than those of more re-
mainly determined by the equivalence ratio, whereas the ODW surface active fuels, such as hydrogen. Experiments on the ignition of liquid
position was insensitive to the variation of the equivalence ratio. Ren fuel in supersonic flows require a shock tube or wind tunnel with
et al. [170] numerically studied the oblique detonation waves in two- high inflow enthalpy and the study of a two-phase shock tube re-
phase kerosene–air mixtures over a wedge for the first time. They found mains highly challenging.
that as the mass flow rate of droplets increases, a shift from a smooth (2) In the study of supersonic atomization, it is important to understand
transition with a curved shock to an abrupt one was found and the the physical mechanisms that are attributed to the primary and
initiation length increases, associated with the increase of the transition secondary breakups, such as the competing effects of K-H instability
pressure, as shown in Fig. 31. By increasing the initial droplet size, a and R-T instability. Experimental techniques must be developed for
smooth transition pattern was observed. The initiation length as well as determining the droplet properties, such as the droplet size and
the transition pressure depends on the spray equivalence ratio, and it is velocity. For numerical studies, no efficient tool for simulating the
mainly owing to the interplay between the evaporative cooling and entire process from the primary to the secondary breakup is avail-
chemical heat release [171]. In general, the research approaches for the able. In addition, more efforts need to be expended on investigating
shock-induced detonation are mainly restricted to the numerical the interaction of strong shock waves (Ms > 3) with droplets.

Fig. 29. Pressure distributions of the oblique detonation


structure: abrupt transition and smooth transition [162].

16
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

Fig. 30. Schematic picture of V-shaped


flame (a) and V + Y Mach stem [168].

Fig. 31. Distributions of dimensionless pressure P/P0 (upper panel) and dimensionless temperature T/T0 (lower panel) for gaseous fuel (left), mixture of gaseous and
liquid fuels (middle), and liquid fuel (right) [170].

(3) The fluid compressibility in supersonic flows results in strong 4. Concluding remarks
compression (eddy shocklets) and dilatation, which affect the dis-
persion and evaporation of droplets. In turn, the dispersed phase In this paper, the physics of supersonic spray combustion were re-
influences the carrier phase due to the interphase interaction. viewed with a focus on applications in scramjet engines, and the status,
Previous research has demonstrated the segregation of inertial progress and challenges of research in this area were discussed.
particles in a different way in compressible turbulence. However, Supersonic spray combustion includes complex physiochemical
the effects of flow compressibility on evaporation and two-phase processes, such as atomization, evaporation, mixing between fuel and
mixing are not well understood in supersonic flows. The interac- air, ignition and flame propagation. In addition, complex wave struc-
tions between the dispersed droplets and the carrier flow with high tures and compressible turbulence occur in the combustion chamber of
compressibility require further investigations. scramjet engines. The interactions among turbulence, spray, shock
(4) Shock waves in the supersonic combustion scramjet interact with waves and combustion inhibit the understanding of the physics of su-
local turbulence and flames. Shock waves can not only accelerate personic spray combustion. Although advances have been made in both
the mixing via the introduction of baroclinic torque into the flow of these research fields, they have yet to be translated into a cohesive
field but also enhance combustion due to the increase of post-shock approach for computing the full structure of sprays combustion, from
pressure/temperature. Most previous research focused on the gas- the atomization of a liquid jet to the formation, evaporation and com-
eous fuel. The interactions among shock waves, fuel spray and bustion of the dispersed fuel droplets. While outstanding issues remain,
chemical reactions and the shock-induced supersonic deflagration/ particularly with respect to the spray combustion in supersonic flows, it
detonation of liquid fuel are worthy of detailed analysis to improve is important to bring together the subsonic and supersonic spray com-
the performance of combustion in scramjets. munities to make progress in the development of overall modeling
(5) The supersonic combustion dynamics in struts or cavities and their approaches. More detailed and advanced diagnostics and experimental
coupling with the shear layer and shock train oscillations in a su- techniques must be developed to enable measurements of spray com-
personic combustor must be studied. This topic is well reviewed in bustion fields in supersonic flows. Therefore, substantial research on
this paper [175]. Readers can refer to the progress that has been fundamental mechanisms and engineering applications is necessary.
made in this area and the solutions that have been proposed.
(6) Other research topics that are not discussed in this review but are Acknowledgements
important for understanding supersonic spray combustion physics
include the following: the development of numerical models for The corresponding author thanks the partial support from the
supersonic spray combustion, accounting for multiscale interactions Nature Science Foundation of China under the grant No. 51676111 and
between gaseous and liquid phases in compressible turbulence; a NSAF under the grant No. U1730104. The first author thanks the partial
reduced chemical reaction mechanism for predicting the ignition, support from the Nature Science Foundation of China under the grant
flame speed, and equilibrium flame temperature of heavy liquid No. 51806179.
hydrocarbon fuels (although the topic of chemical reaction me-
chanisms is critical for supersonic spray combustion, it is outside Appendix A. Supplementary data
the scope of this review); an experimental database of supersonic
spray combustion for high-resolution simulations to demonstrate Supplementary data to this article can be found online at https://
the accuracy and the capability of obtaining such physical results; doi.org/10.1016/j.paerosci.2018.12.002.
and integration of numerical models into simulations for verifica-
tion, experimental validation, and predictive engineering analysis References
of scramjet engine performance under uncertainty.
[1] J.M. Tishkoff, J.P. Drummond, T. Edwards, et al., Future directions of supersonic
combustion research - air Force/NASA workshop on supersonic combustion, AIAA
Paper (1997) 97–1017.

17
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

[2] A. Ben-Yakar, R.K. Hanson, Cavity flame-holders for ignition and flame stabili- [39] G.L. Brown, A. Roshko, On density effects and large structure in turbulent mixing
zation in scramjets: an overview, J. Propul. Power 17 (4) (2001) 869–877. layers, J. Fluid Mech. 64 (04) (1974) 775–816.
[3] A. Ferri, Review of problems in application of supersonic combustion, J. Roy. [40] M. Samimy, G.S. Elliott, Effects of compressibility on the characteristics of free
Aeronaut. Soc. 68 (1964) 575–597. shear layers, AIAA J. 28 (3) (1990) 439–445.
[4] A.C. Idris, M.R. Saad, H. Zare-Behtash, et al., Luminescent measurement systems [41] G.S. Elliott, M. Samimy, Compressibility effects in free shear layers, Phys. Fluid.
for the investigation of a scramjet inlet-isolator, Sensors 14 (4) (2014) 6606–6632. Fluid Dynam. 2 (7) (1990) 1231–1240.
[5] R.S. Fry, A century of ramjet propulsion technology evolution, J. Propul. Power 20 [42] J.P. Bonnet, J.R. Debisschop, Experimental Studies of uhe Turbulent Structure of
(1) (2004) 27–58. Supersonic Mixing Layers, (1993) AIAA-93-0217.
[6] J. Urzay, Supersonic combustion in air-breathing propulsion systems for hy- [43] S. Barre, C. Quine, J.P. Dussauge, Compressibility effects on the structure of su-
personic flight, Annu. Rev. Fluid Mech. 50 (2018) 593–627. personic mixing layers: experimental results, J. Fluid Mech. 259 (4) (1994) 47–78.
[7] W. Huang, Z. Du, L. Yan, et al., Flame propagation and stabilization in dual-mode [44] N.T. Clemens, M.G. Mungal, Large-scale structure and entrainment in the super-
scramjet combustors: a survey, Prog. Aero. Sci. 101 (2018) 13–30. sonic mixing layer, J. Fluid Mech. 284 (1995) 171–216.
[8] D. Martínez-Ruiz, J. Urzay, A.L. Sánchez, et al., Dynamics of thermal ignition of [45] A.W. Vreman, N.D. Sandham, K.H. Luo, Compressible mixing layer growth rate
spray flames in mixing layers, J. Fluid Mech. 734 (2013) 387–423. and turbulence characteristics, J. Fluid Mech. 320 (1996) 235–258.
[9] A. Chaudhuri, A. Hadjadj, A. Chinnayya, et al., Numerical study of compressible [46] C. Pantano, S. Sarkar, A study of compressibility effects in the high-speed turbu-
mixing layers using high-order WENO schemes, J. Sci. Comput. 47 (2) (2011) lent shear layer using direct simulation, J. Fluid Mech. 451 (2002) 329–371.
170–197. [47] A. Atoufi, M. Fathali, B. Lessani, Compressibility effects and turbulent kinetic
[10] T.F. Fric, A. Roshko, Vortical structure in the wake of a transverse jet, J. Fluid energy exchange in temporal mixing layers, J. Turbul. 16 (7) (2015) 676–703.
Mech. 279 (1994) 1–47. [48] B. Wang, W. Wei, Y. Zhang, et al., Passive scalar mixing in Mc < 1 planar shear
[11] K. Manoharan, S. Hemchandra, Absolute/convective instability transition in a layer flows, Comput. Fluid 123 (2015) 32–43.
backward facing step combustor: fundamental mechanism and influence of density [49] P.J.M. Ferrer, G. Lehnasch, A. Mura, Compressibility and heat release effects in
gradient, J. Eng. Gas Turbines Power 137 (2) (2015) 021501. high-speed reactive mixing layers I. Growth rates and turbulence characteristics,
[12] F.W. Barnes, C. Segal, Cavity-based flameholding for chemically-reacting super- Combust. Flame 180 (2016) 284–303.
sonic flows, Prog. Aero. Sci. 76 (2015) 24–41. [50] J.B. Perurena, C.O. Asma, R. Theunissen, et al., Experimental investigation of li-
[13] P. Jenny, D. Roekaerts, N. Beishuizen, Modeling of turbulent dilute spray com- quid jet injection into Mach 6 hypersonic crossflow, Exp. Fluid 46 (3) (2009)
bustion, Prog. Energy Combust. Sci. 38 (6) (2012) 846–887. 403–417.
[14] X. Jiang, G.A. Siamas, K. Jagus, et al., Physical modelling and advanced simula- [51] F. Xiao, Z.G. Wang, M.B. Sun, et al., Simulation of drop deformation and breakup
tions of gas–liquid two-phase jet flows in atomization and sprays, Prog. Energy in supersonic flow, Proc. Combust. Inst. 36 (2) (2017) 2417–2424.
Combust. Sci. 36 (2) (2010) 131–167. [52] D.B. Bogy, Drop formation in a circular liquid jet, Annu. Rev. Fluid Mech. 11 (1)
[15] M. Broumand, M. Birouk, Liquid jet in a subsonic gaseous crossflow: recent pro- (1979) 207–228.
gress and remaining challenges, Prog. Energy Combust. Sci. 57 (2016) 1–29. [53] J.C. Lasheras, E.J. Hopfinger, Liquid jet instability and atomization in a coaxial gas
[16] S. Balachandar, J.K. Eaton, Turbulent dispersed multiphase flow, Annu. Rev. Fluid stream, Annu. Rev. Fluid Mech. 32 (1) (2000) 275–308.
Mech. 42 (2010) 111–133. [54] X. Chen, D. Ma, V. Yang, S. Popinet, High-fidelity simulations of impinging jet
[17] A.R. Masri, Turbulent combustion of sprays: from dilute to dense, Combust. Sci. atomization, Atomization Sprays 23 (2013) 1079–1101.
Technol. 188 (10) (2016) 1619–1639. [55] Q. Fu, L. Yang, K. Cui, F. Zhuang, Effects of orifice geometry on gelled propellants
[18] Y. Yang, J. Wang, Y. Shi, et al., Interactions between inertial particles and sprayed from impinging-jet injectors, J. Propul. Power 30 (2014) 1113–1117.
shocklets in compressible turbulent flow, Phys. Fluid. 26 (9) (2014) 091702. [56] N.S. Rodrigues, V. Kulkarni, J. Gao, J. Chen, P.E. Sojka, An experimental and
[19] Q. Zhang, H. Liu, Z. Ma, et al., Preferential concentration of heavy particles in theoretical investigation of spray characteristics of impinging jets in impact wave
compressible isotropic turbulence, Phys. Fluid. 28 (5) (2016) 055104. regime, Exp. Fluid 56 (2015) 50.
[20] Q. Dai, K. Luo, T. Jin, et al., Direct numerical simulation of turbulence modulation [57] P. Zhang, B. Wang, Effects of elevated ambient pressure on the disintegration of
by particles in compressible isotropic turbulence, J. Fluid Mech. 832 (2017) impinged sheets, Phys. Fluids 29 (4) (2017) 042102.
438–482. [58] A. Sherman, J. Schetz, Breakup of liquid sheets and jets in a supersonic gas stream,
[21] A. Ingenito, C. Bruno, Physics and regimes of supersonic combustion, AIAA J. 48 AIAA J. 9 (4) (1971) 666–673.
(3) (2010) 515–525. [59] R.H. Thomas, J.A. Schetz, Distributions across the plume of transverse liquid and
[22] L. Biagioni, L. d'Agostino, Measurement of Energy Spectra in Weakly Compressible slurry jets in supersonic airflow, AIAA J. 23 (12) (1985) 1892–1901.
Turbulence, 30th AIAA Fluid Dynamics Conference, Norfolk, VA, 1999. [60] P.K. Wu, K.A. Kirkendall, R.P. Fuller, Breakup processes of liquid jets in subsonic
[23] C. Segal, The Scramjet Engine: Processes and Characteristics, Cambridge crossflows, J. Propul. Power 13 (1) (1997) 64–73.
University Press, 2009. [61] P.K. Wu, K.A. Kirkendall, R.P. Fuller, Spray structures of liquid jets atomized in
[24] G. Balakrishnan, F.A. Williams, Turbulent combustion regimes for hypersonic subsonic crossflows, J. Propul. Power 14 (2) (1998) 173–182.
propulsion employing hydrogen-air diffusion flames, J. Propul. Power 10 (3) [62] D.S. Olinger, K.A. Sallam, Digital holographic analysis of the near field of aerated-
(1994) 434–437. liquid jets in crossflow, J. Propul. Power 30 (6) (2014) 1636–1645.
[25] M.B. Colket, L.J. Spadaccini, Scramjet fuels autoignition study, J. Propul. Power [63] D. Sedarsky, M. Paciaroni, E. Berrocal, et al., Model validation image data for
17 (2) (2001) 315–323. breakup of a liquid jet in crossflow: part I, Exp. Fluid 49 (2) (2010) 391–408.
[26] T. Mitani, N. Chinzei, T. Kanda, Reaction and mixing-controlled combustion in [64] J. Song, K. Ahn, M. Kim, et al., Effects of orifice internal flow on liquid jets in
scramjet engines, J. Propul. Power 17 (2) (2001) 308–314. subsonic crossflows, J. Propul. Power 27 (3) (2011) 608–619.
[27] J.B. Freund, S.K. Lele, P. Moin, Compressibility effects in a turbulent annular [65] K.C. Lin, P.J. Kennedy, T.A. Jackson, Structures of water jets in a Mach 1.94 su-
mixing layer. Part 1. Turbulence and growth rate, J. Fluid Mech. 421 (2000) personic crossflow, AIAA Paper (2004) 0971.
229–267. [66] T. Regert, I. Horvath, J.M. Buchlin, et al., Study on breakup of liquid ligaments in
[28] J.B. Freund, P. Moin, S.K. Lele, Compressibility effects in a turbulent annular hypersonic cross flow using laser sheet imaging and infrared light extinction
mixing layer. Part 2. Mixing of a passive scalar, J. Fluid Mech. 421 (2000) spectroscopy, Prog. Flight Phys. 9 (9) (2017) 229–250.
269–292. [67] K.C. Lin, M.C. Lai, T. Ombrello, et al., Structures and temporal evolution of liquid
[29] S. Kida, S.A. Orszag, Energy and spectral dynamics in forced compressible tur- jets in supersonic crossflow, 55th AIAA Aerospace Sciences Meeting. 2017, 1958.
bulence, J. Sci. Comput. 5 (2) (1990) 85–125. [68] K.A. Sallam, K.C. Lin, S.D. Hammack, et al., Digital holographic analysis of the
[30] R. Jahanbakhshi, C.K. Madnia, Entrainment in a compressible turbulent shear breakup of aerated liquid jets in supersonic crossflow, 55th AIAA Aerospace
layer, J. Fluid Mech. 797 (2016) 564–603. Sciences Meeting. 2017, 1957.
[31] J. O'Brien, J. Urzay, M. Ihme, et al., Subgrid-scale backscatter in reacting and inert [69] F. Xiao, Z.G. Wang, M.B. Sun, et al., Large eddy simulation of liquid jet primary
supersonic hydrogen–air turbulent mixing layers, J. Fluid Mech. 743 (2014) breakup in supersonic air crossflow, Int. J. Multiphas. Flow 87 (2016) 229–240.
554–584. [70] X. Li, M.C. Soteriou, M. Arienti, et al., High-fidelity simulation of atomization and
[32] J.Y. Choi, F. Ma, V. Yang, Combustion oscillations in a scramjet engine combustor evaporation in a liquid jet in cross-flow, AIAA Paper (2011-0099) 1–19.
with transverse fuel injection, Proc. Combust. Inst. 30 (2) (2005) 2851–2858. [71] M. Rachner, J. Becker, C. Hassa, et al., Modelling of the atomization of a plain
[33] Z. Ren, B. Wang, L. Zheng, Numerical analysis on interactions of vortex, shock liquid fuel jet in crossflow at gas turbine conditions, Aero. Sci. Technol. 6 (7)
wave, and exothermal reaction in a supersonic planar shear layer laden with (2002) 495–506.
droplets, Phys. Fluids 30 (3) (2018) 036101. [72] A. Mashayek, M. Behzad, N. Ashgriz, Multiple injector model for primary breakup
[34] Z. Ren, B. Wang, D. Zhao, et al., Flame propagation involved in vortices of su- of a liquid jet in crossflow, AIAA J. 49 (11) (2011) 2407–2420.
personic mixing layers laden with droplets: effects of ambient pressure and spray [73] H.T. Chang, L.W. Hourng, L.C. Chien, Application of flux-vector-splitting scheme
equivalence ratio, Phys. Fluids 30 (10) (2018) 106107. to a dilute gas-particle JPL nozzle flow, Int. J. Numer. Methods Fluid. 22 (1996)
[35] Z. Ren, B. Wang, L. Zheng, et al., Numerical studies on supersonic spray com- 921–935.
bustion in high-temperature shear flows in a scramjet combustor, Chin. J. [74] T.G. Theofanous, Aerobreakup of Newtonian and viscoelastic liquids, Annu. Rev.
Aeronaut. 31 (9) (2018) 1870–1879. Fluid Mech. 43 (2011) 661–690.
[36] P.J.M. Ferrer, R. Buttay, G. Lehnasch, et al., A detailed verification procedure for [75] A.R. Hanson, E.G. Domich, H.S. Adams, Shock tube investigation of the breakup of
compressible reactive multicomponent Navier–Stokes solvers, Comput. Fluid 89 drops by air blasts, Phys. Fluid. 6 (8) (1963) 1070–1080.
(2014) 88–110. [76] M. Pilch, C.A. Erdman, Use of breakup time data and velocity history data to
[37] M.D. Slessor, M. Zhuang, P.E. Dimotakis, Turbulent shear-layer mixing: growth- predict the maximum size of stable fragments for acceleration-induced breakup of
rate compressibility scaling, J. Fluid Mech. 414 (2000) 35–45. a liquid drop, Int. J. Multiphas. Flow 13 (6) (1987) 741–757.
[38] J.R. Nuding, Interaction of compressible shear layers with shock waves: an ex- [77] D.D. Joseph, J. Belanger, G.S. Beavers, Breakup of a liquid drop suddenly exposed
perimental study, Part I, AIAA Paper (1996) 4515. to a high-speed airstream, Int. J. Multiphas. Flow 25 (6–7) (1999) 1263–1303.

18
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

[78] L.P. Hsiang, G. FAETH, Secondary drop breakup in the deformation regime, 30th [113] K. Kumaran, V. Babu, Mixing and combustion characteristics of kerosene in a
Aerospace Sciences Meeting and Exhibit, 1992, p. 110. model supersonic combustor, J. Propul. Power 25 (3) (2009) 583–592.
[79] L.P. Hsiang, G. Faeth, Deformation and secondary breakup of drops, 31st [114] Y. Pan, J.G. Tan, J.H. Liang, et al., Experimental investigation of combustion
Aerospace Sciences Meeting, 1992, p. 814. mechanisms of kerosene-fueled scramjet engines with double-cavity flameholders,
[80] L.P. Hsiang, G.M. Faeth, Drop deformation and breakup due to shock wave and Acta Mech. Sin. 27 (6) (2011) 891–897.
steady disturbance, Previews Heat Mass Transf. 5 (21) (1995) 415. [115] M. Tsue, O. Imamura, S. Suzuki, et al., Effects of fuel properties on self-ignition
[81] L.P. Hsiang, G.M. Faeth, Drop deformation and breakup due to shock wave and and flame-holding performances in SCRAM jet combustor for n-alkane fuels, Proc.
steady disturbances, Int. J. Multiphas. Flow 21 (4) (1995) 545–560. Combust. Inst. 33 (2) (2011) 2391–2398.
[82] D. Igra, K. Takayama, Investigation of aerodynamic breakup of a cylindrical water [116] H. Bao, J. Zhou, Y. Pan, Effect of cavity configuration on kerosene spark ignition in
droplet, Atomization Sprays 11 (2) (2001). a scramjet combustor at Ma 4.5 flight condition, Acta Astronaut. 117 (2015)
[83] D. Igra, K. Takayama, Numerical simulation of shock wave interaction with a 368–375.
water column, Shock Waves 11 (3) (2001) 219–228. [117] L. Gang, Z. Shaohua, T. Liang, et al., Numerical investigation of the effect of re-
[84] S. Sembian, M. Liverts, N. Tillmark, et al., Plane shock wave interaction with a action models on the supersonic combustion of liquid kerosene, 51st AIAA/SAE/
cylindrical water column, Phys. Fluids 28 (5) (2016) 056102. ASEE Joint Propulsion Conference, 2015, p. 4167.
[85] C. Aalburg, B. Leer, G.M. Faeth, Deformation and drag properties of round drops [118] Z.P. Wang, F. Li, H.B. Gu, et al., Experimental study on the effect of combustor
subjected to shock-wave disturbances, AIAA J. 41 (12) (2003) 2371–2378. configuration on the performance of dual-mode combustor, Aero. Sci. Technol. 42
[86] J.C. Meng, T. Colonius, Numerical simulations of the early stages of high-speed (2015) 169–175.
droplet breakup, Shock Waves 25 (4) (2015) 399–414. [119] Y. Zong, W. Bao, J. Chang, et al., Effect of fuel injection allocation on the com-
[87] G. Xiang, B. Wang, Numerical study of a planar shock interacting with a cylind- bustion characteristics of a cavity-strut model scramjet, J. Aero. Eng. 28 (1) (2013)
rical water column embedded with an air cavity, J. Fluid Mech. 825 (2017) 04014050.
825–852. [120] H. Bao, J. Zhou, Y. Pan, et al., Spark ignition of liquid kerosene in scramjet
[88] N. Liu, Z. Wang, M. Sun, et al., Numerical simulation of liquid droplet breakup in combustor equipped with partly-covered cavity, J. Propul. Power 31 (4) (2015)
supersonic flows, Acta Astronaut. 145 (2018) 116–130. 1014–1018.
[89] J.C. Meng, T. Colonius, Numerical simulation of the aerobreakup of a water [121] S. Nakaya, Y. Hikichi, Y. Nakazawa, et al., Ignition and supersonic combustion
droplet, J. Fluid Mech. 835 (2018) 1108–1135. behavior of liquid ethanol in a scramjet model combustor with cavity flame
[90] J.C. Hermanson, P.E. Dimotakis, Effects of heat release in a turbulent, reacting holder, Proc. Combust. Inst. 35 (2) (2015) 2091–2099.
shear layer, J. Fluid Mech. 199 (1989) 333–375. [122] M. Sun, Z. Zhong, J. Liang, et al., Experimental investigation on combustion
[91] R.S. Barlow, D.C. Fourguette, M.G. Mungal, et al., Experiments on the structure of performance of cavity-strut injection of supercritical kerosene in supersonic model
an annular compressible reacting shear layer, AIAA J. 30 (9) (1992) 2244–2251. combustor, Acta Astronaut. 127 (2016) 112–119.
[92] M.F. Miller, C.T. Bowman, M.G. Mungal, An experimental investigation of the [123] J.M. Seiner, S.M. Dash, D.C. Kenzakowski, Historical survey on enhanced mixing
effects of compressibility on a turbulent reacting mixing layer, J. Fluid Mech. 356 in scramjet engines, J. Propul. Power 17 (6) (2001) 1273–1286.
(1998) 25–64. [124] A. Kumar, D. Bushnell, M. Hussaini, A mixing augmentation technique for hy-
[93] W.H. Calhoon, S. Arunajatesan, S.M. Dash, Heat release and compressibility effects pervelocity scramjets, J. Propul. Power 5 (5) (1987) 514–522.
on planar shear layer development, AIAA Paper 1273 (2003) 2003. [125] F.E. Marble, Gasdynamic enhancement of nonpremixed combustion, Symposium
[94] I. Mahle, H. Foysi, S. Sarkar, et al., On the turbulence structure in inert and re- (International) on Combustion, vol 25, Elsevier, 1994, pp. 1–12 1.
acting compressible mixing layers, J. Fluid Mech. 593 (2007) 171–180. [126] H. Huh, J.F. Driscoll, Shock-wave-enhancement of the mixing and the stability
[95] J. Mathew, I. Mahle, R. Friedrich, Effects of compressibility and heat release on limits of supersonic hydrogen-air jet flames, Symposium (International) on
entrainment processes in mixing layers, J. Turbul. (9) (2008) N14. Combustion, vol 26, Elsevier, 1996, pp. 2933–2939 2.
[96] P.J.M. Ferrer, G. Lehnasch, A. Mura, Compressibility and heat release effects in [127] H. Huh, J. Driscoll, Measured effects of shock waves on supersonic hydrogen-air
high-speed reactive mixing layers II. Structure of the stabilization zone and flames, 32nd Joint Propulsion Conference and Exhibit, 1996, p. 3035.
modeling issues relevant to turbulent combustion in supersonic flows, Combust. [128] H. Huh, J. Kim, J. Driscoll, Measured characteristics of flow and combustion in
Flame 180 (2016) 304–320. supersonic flame/shock wave interaction, 37th Joint Propulsion Conference and
[97] B. Wang, W. Wei, S. Ma, et al., Construction of one-step H2/O2 reaction me- Exhibit, 2001, p. 3935.
chanism for predicting ignition and its application in simulation of supersonic [129] Z. Li, F. Jaberi, Numerical investigations of shock-turbulence interaction in a
combustion, Int. J. Hydrogen Energy 41 (42) (2016) 19191–19206. planar mixing layer, 48th AIAA Aerospace Sciences Meeting Including the New
[98] Y.L. Zhang, B. Wang, H.Q. Zhang, Ignition, flame propagation and extinction in Horizons Forum and Aerospace Exposition, 2010, p. 112.
the supersonic mixing layer flow, Sci. China Technol. Sci. 57 (11) (2014) [130] J.H. Kim, Y. Yoon, I.S. Jeung, et al., Numerical study of mixing enhancement by
2256–2264. shock waves in model scramjet engine, AIAA J. 41 (6) (2003) 1074–1080.
[99] C. Qian, W. Bing, Z. Huiqiang, et al., Numerical investigation of H2/air combus- [131] F. Génin, S. Menon, Studies of shock/turbulent shear layer interaction using large-
tion instability driven by large scale vortex in supersonic mixing layers, Int. J. eddy simulation, Comput. Fluid 39 (5) (2010) 800–819.
Hydrogen Energy 41 (4) (2016) 3171–3184. [132] Y. Zhang, B. Wang, H. Zhang, et al., Mixing enhancement of compressible planar
[100] Z. Ren, B. Wang, S. Yang, et al., Evolution of flame kernel in one eddy turnover of mixing layer impinged by oblique shock waves, J. Propul. Power 31 (1) (2014)
high-speed droplet laden shear layers, J. Loss Prev. Process. Ind. 49 (2017) 156–169.
938–946. [133] S.H. Lee, I.S. Jeung, Y. Yoon, Computational investigation of shock-enhanced
[101] Z. Ren, B. Wang, Q. Xie, et al., Thermal auto-ignition in high-speed droplet-laden mixing and combustion, AIAA J. 35 (12) (1997) 1813–1820.
mixing layers, Fuel 191 (2017) 176–189. [134] C.J. Roy, J.R. Edwards, Numerical simulation of a three-dimensional flame/shock
[102] O.A. Powell, J.T. Edwards, R.B. Norris, et al., Development of hydrocarbon-fueled wave interaction, AIAA J. 38 (5) (2000) 745–754.
scramjet engines: the hypersonic technology (HyTech) program, J. Propul. Power [135] A. Ratner, J.F. Driscoll, H. Huh, et al., Combustion efficiencies of supersonic
17 (6) (2001) 1170–1176. flames, J. Propul. Power 17 (2) (2001) 301–307.
[103] P.H. Renard, D. Thevenin, J.C. Rolon, et al., Dynamics of flame/vortex interac- [136] H. Nakamura, N. Sato, H. Kobayashi, et al., Effect of the location of an incident
tions, Prog. Energy Combust. Sci. 26 (3) (2000) 225–282. shock wave on combustion and flow field of wall fuel-injection, Trans. Jpn. Soc.
[104] H. Miyajima, N. Chinzei, T. Mitani, et al., Development status of the NAL ramjet Aeronaut. Space Sci. 51 (173) (2008) 170–175.
engine test facility and sub-scale scramjet engine, AIAA 4th International [137] T. Mai, Y. Sakimitsu, H. Nakamura, et al., Effect of the incident shock wave in-
Aerospace Planes Conference, AIAA, 1992 92-5094. teracting with transversal jet flow on the mixing and combustion, Proc. Combust.
[105] P. Manna, R. Behera, D. Chakraborty, Liquid-fueled strut-based scramjet com- Inst. 33 (2) (2011) 2335–2342.
bustor design: a computational fluid dynamics approach, J. Propul. Power 24 (2) [138] A.A. Shekarian, S. Tabejamaat, Y. Shoraka, Effects of incident shock wave on
(2008) 274–281. mixing and flame holding of hydrogen in supersonic air flow, Int. J. Hydrogen
[106] K. Kumaran, P.R. Behera, V. Babu, Numerical investigation of the supersonic Energy 39 (19) (2014) 10284–10292.
combustion of kerosene in a strut-based combustor, J. Propul. Power 26 (5) (2010) [139] V. Hariharan, R.K. Velamati, C. Prathap, Investigation on supersonic combustion
1084–1091. of hydrogen with variation of combustor inlet conditions, Int. J. Hydrogen Energy
[107] V.A. Vinogradov, S.A. Kobigskij, M.D. Petrov, Experimental investigation of ker- 41 (13) (2016) 5833–5841.
osene fuel combustion in supersonic flow, J. Propul. Power 11 (1) (1995) [140] C. Huete, A.L. Sánchez, F.A. Williams, et al., Diffusion-flame ignition by shock-
130–134. wave impingement on a supersonic mixing layer, J. Fluid Mech. 784 (2015)
[108] T. Kouchi, G. Masuya, T. Mitani, et al., Mechanism and control of combustion- 74–108.
mode transition in a scramjet engine, J. Propul. Power 28 (1) (2012) 106–112. [141] C. Huete, A.L. Sánchez, F.A. Williams, Diffusion-flame ignition by shock-wave
[109] C. Zhang, Q. Yang, J. Chang, et al., Nonlinear characteristics and detection of impingement on a hydrogen–air supersonic mixing layer, J. Propul. Power
combustion modes for a hydrocarbon fueled scramjet, Acta Astronaut. 110 (2015) (2016) 1–8.
89–98. [142] P.M. Rubins, R.C. Bauer, Review of shock-induced supersonic combustion research
[110] Q. Yang, K. Chetehouna, N. Gascoin, et al., Experimental study on combustion and hypersonic applications, J. Propul. Power 10 (5) (1994) 593–601.
modes and thrust performance of a staged-combustor of the scramjet with dual- [143] J. Yang, T. Kubota, E.E. Zukoski, Applications of shock-induced mixing to super-
strut, Acta Astronaut. 122 (2016) 28–34. sonic combustion, AIAA J. 31 (5) (1993) 854–862.
[111] G. Yu, J.G. Li, X.Y. Chang, et al., Fuel injection and flame stabilization in a liquid- [144] J.A. Nicholls, E.K. Dabora, An Experimental and Theoretical Study of Stationary
kerosene-fueled supersonic combustor, J. Propul. Power 19 (5) (2003) 885–893. Gaseous Detonation Waves, MICHIGAN UNIV ANN ARBOR OFFICE OF
[112] G. Yu, J.G. Li, J.R. Zhao, et al., An experimental study of kerosene combustion in a RESEARCH ADMINISTRATION, 1961.
supersonic model combustor using effervescent atomization, Proc. Combust. Inst. [145] R.P. Rhodes, P.M. Rubins, Shock-induced combustion with oblique shocks-com-
30 (2) (2005) 2859–2866. parison of experiment and kinetic calculations, AIAA J. 1 (12) (1963) 2778–2784.

19
Z. Ren et al. Progress in Aerospace Sciences xxx (xxxx) xxx–xxx

[146] J. Kasahara, T. Fujiwara, T. Endo, Chapman-Jouguet oblique detonation structure to propulsion, J. Propul. Power 22 (6) (2006) 1230.
around hypersonic projectiles, AIAA J. 39 (8) (2001) 1553–1561. [162] H.H. Teng, Z.L. Jiang, On the transition pattern of the oblique detonation struc-
[147] J. Kasahara, T. Arai, S. Chiba, et al., Criticality for stabilized oblique detonation ture, J. Fluid Mech. 713 (2012) 659–669.
waves around spherical bodies in acetylene/oxygen/krypton mixtures, Proc. [163] H.H. Teng, Z.L. Jiang, H.D. Ng, Numerical study on unstable surfaces of oblique
Combust. Inst. 29 (2) (2002) 2817–2824. detonations, J. Fluid Mech. 744 (2014) 111–128.
[148] S. Maeda, R. Inada, J. Kasahara, et al., Visualization of the non-steady state ob- [164] H. Teng, Y. Zhang, Z. Jiang, Numerical investigation on the induction zone
lique detonation wave phenomena around hypersonic spherical projectile, Proc. structure of the oblique detonation waves, Comput. Fluid 95 (2014) 127–131.
Combust. Inst. 33 (2) (2011) 2343–2349. [165] T. Wang, Y. Zhang, H. Teng, et al., Numerical study of oblique detonation wave
[149] S. Maeda, J. Kasahara, A. Matsuo, Oblique detonation wave stability around a initiation in a stoichiometric hydrogen-air mixture, Phys. Fluids 27 (9) (2015)
spherical projectile by a high time resolution optical observation, Combust. Flame 096101.
159 (2) (2012) 887–896. [166] H. Teng, H.D. Ng, Z. Jiang, Initiation Characteristics of Wedge-induced Oblique
[150] S. Maeda, S. Sumiya, J. Kasahara, et al., Initiation and sustaining mechanisms of Detonation Waves in a Stoichiometric Hydrogen-air Mixture, Proceedings of the
stabilized oblique detonation waves around projectiles, Proc. Combust. Inst. 34 (2) Combustion Institute, 2016.
(2013) 1973–1980. [167] Y. Zhang, J. Gong, T. Wang, Numerical study on initiation of oblique detonations
[151] S. Maeda, S. Sumiya, J. Kasahara, et al., Scale effect of spherical projectiles for in hydrogen–air mixtures with various equivalence ratios, Aero. Sci. Technol. 49
stabilization of oblique detonation waves, Shock Waves 25 (2) (2015) 141–150. (2016) 130–134.
[152] K. Iwata, S. Nakaya, M. Tsue, Numerical investigation of the effects of nonuniform [168] K. Iwata, S. Nakaya, M. Tsue, Wedge-stabilized oblique detonation in an in-
premixing on shock-induced combustion, AIAA J. 54 (2) (2016) 1682–1692. homogeneous hydrogen–air mixture, Proc. Combust. Inst. 36 (2) (2017)
[153] J.L. Cambier, H. Adelman, G.P. Menees, Numerical simulations of an oblique 2761–2769.
detonation wave engine, J. Propul. Power 6 (3) (1990) 315–323. [169] Y. Fang, Z. Hu, H. Teng, et al., Numerical study of inflow equivalence ratio in-
[154] D.T. Pratt, J.W. Humphrey, D.E. Glenn, Morphology of standing oblique detona- homogeneity on oblique detonation formation in hydrogen–air mixtures, Aero.
tion waves, J. Propul. Power 7 (5) (1991) 837–845. Sci. Technol. 71 (2017) 256–263.
[155] C. Li, K. Kailasanath, E.S. Oran, Detonation structures behind oblique shocks, [170] Z. Ren, B. Wang, G. Xiang, et al., Effect of the multiphase composition in a pre-
Phys. Fluid. 6 (4) (1994) 1600–1611. mixed fuel–air stream on wedge-induced oblique detonation stabilisation, J. Fluid
[156] A.A. Thaker, H.K. Chelliah, Numerical Prediction of Oblique Detonation Wave Mech. 846 (2018) 411–427.
Structures Using Detailed and Reduced Reaction Mechanisms, (1997). [171] Z. Ren, B. Wang, G. Xiang, et al., Numerical Analysis of Wedge-induced Oblique
[157] J.Y. Choi, D.W. Kim, I.S. Jeung, et al., Cell-like structure of unstable oblique de- Detonations in Two-phase Kerosene–air Mixtures, Proceedings of the Combustion
tonation wave from high-resolution numerical simulation, Proc. Combust. Inst. 31 Institute, 2018.
(2) (2007) 2473–2480. [172] Y. Ju, W. Sun, Plasma assisted combustion: dynamics and chemistry, Prog. Energy
[158] J. Verreault, A.J. Higgins, R.A. Stowe, Formation and structure of steady oblique Combust. Sci. 48 (2015) 21–83.
and conical detonation waves, AIAA J. 50 (8) (2012) 1766–1772. [173] B.R. Halls, J.R. Gord, N. Jiang, et al., High-speed three-dimensional tomographic
[159] M. Gui, B. Fan, Wavelet structure of wedge-induced oblique detonation waves, measurements for combustion systems, 32nd AIAA Aerodynamic Measurement
Combust. Sci. Technol. 184 (10–11) (2012) 1456–1470. Technology and Ground Testing Conference, 2016, p. 4027.
[160] C. Viguier, L.F.F. da Silva, D. Desbordes, et al., Onset of oblique detonation waves: [174] E.D. Gonzalez-Juez, A.R. Kerstein, R. Ranjan, et al., Advances and challenges in
comparison between experimental and numerical results for hydrogen-air mix- modeling high-speed turbulent combustion in propulsion systems, Prog. Energy
tures, Symposium (International) on Combustion, vol 26, Elsevier, 1996, pp. Combust. Sci. 60 (2017) 26–67.
3023–3031 2. [175] J. Chang, J. Zhang, W. Bao, et al., Research progress on strut-equipped supersonic
[161] D.S. Stewart, A.R. Kasimov, State of detonation stability theory and its application combustors for scramjet application, Prog. Aero. Sci. 103 (2018) 1–30.

20

Potrebbero piacerti anche