You are on page 1of 13

JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.

1 (1-13)
Aerospace Science and Technology ••• (••••) •••–•••

1 67
Contents lists available at ScienceDirect
2 68
3 69
4 Aerospace Science and Technology 70
5 71
6 72
7 www.elsevier.com/locate/aescte 73
8 74
9 75
10
Review 76
11 77
12 Transverse jet in supersonic crossflows 78
13 79
14 Wei Huang ∗ 80
15 81
Science and Technology on Scramjet Laboratory, National University of Defense Technology, Changsha, Hunan 410073, People’s Republic of China
16 82
17 83
18 84
a r t i c l e i n f o a b s t r a c t
19 85
20 Article history: 86
The mixing and combustion process plays an important role in the realization of the scramjet engine,
21 Received 14 December 2014 and the transverse injection from a wall orifice is widely employed for the simplest and most promising 87
22 Received in revised form 31 December 2015 of its configurations. In the current survey, the research progress on the transverse jet in supersonic 88
23 Accepted 2 January 2016 89
crossflows has been summarized systematically from four aspects, namely single injection, multiport
Available online xxxx
24 injection, interaction between jet and vortex generator, and interaction between jet and shock wave, 90
25 Keywords: and the basic principle of the transverse injection has been provided as well. At last, some promising 91
26 Aerospace propulsion system recommendations have been proposed, namely the refined vortex structure capture, the mixing and 92
27 Transverse injection combustion process in the novel injector and multiport flow fields, especially with the incident shock 93
28
Supersonic crossflow wave interaction, and the combinatorial operating and optimization process between the fuel injection 94
Inter action mechanism and the vortex generator.
29 95
Shock wave © 2016 Published by Elsevier Masson SAS.
30 Vortex generator 96
31 97
32 98
33 99
34 100
1. Introduction Transverse injection from a wall orifice is one of the sim-
35 101
plest and most promising configurations to enhance the mixing
36 102
The scramjet (supersonic combustion ramjet) engine would be- process between the fuel and air in supersonic flows [3], and it
37 103
come one of the most effective engine cycles for the hypersonic attracts an increasing attention since the early sixties [16], es-
38 104
flight in the near future [1], however, it is very difficult to ob- pecially on some scramjet powered vehicles [17–20], see Fig. 1.
39 105
tain a stable and efficient combustion flow field in a scramjet Fig. 1 represents the computational results obtained by the large
40 106
engine due to the very short residence time of the injectant in- eddy simulation approach for the HyShot II combustor, and the
41 107
side the combustor, namely of the order of milliseconds [2], and mixing process is dominated by the counter–rotating vortex pair
42 108
the mixing process at the molecular level has to be completed in (CVP) and Ω -shaped vortices in the near and far fields respec-
43 109
a limited combustor length [3]. The rapid mixing and combustion tively [21,22], as well as the Kelvin–Helmholtz instabilities induced
44 110
process is crucial for the realization of the scramjet engine, and it by the high levels of upper jet shear layer [23], see Fig. 2. That
45 111
takes place nearly simultaneously in the combustor [4]. The mix- is to say its near-field mixing is predominantly controlled by an
46 112
ing process is the initial phase for all the physical ones, as well entrainment-stretching-mixing process [24], and the large protru-
47 113
as the primary factor to restrict the combustion process, i.e. igni- sions of injectant are induced by the large-scale vortices on the
48 114
49 tion and flame propagation. Increased efficiency in fuel–air mixing windward side of the jet plume [25]. The far-field mixing is con- 115
50 may lead to reductions in size and weight of the engine, as well as trolled by mass diffusion [26]. Horseshoe vortex is obtained by the 116
51 reducing the amount of structure that needs to be cooled [5]. In or- interaction between the incoming boundary layer and the jet, and 117
52 der to promote the mixing process between the injectant and air, it remains close to the wall of injection, wraps around the jet 118
53 many fuel-injection systems have been proposed in recent years, periphery and propagates downstream. Therefore, the horseshoe 119
54 i.e. ramp [6], aerodynamic ramp [7,8], strut [9], pylon [10–12], and vortex does not interact with the jet, and it does not take part 120
55 any other combination, as well as the cantilevered ramp injector in the mixing process [27]. A cavity has been utilized by Lee and 121
56 which has been used as the inlet injection scheme to shorten the Mitani [28] to modify the injector geometry in order to promote 122
57 length of the combustor [13,14]. Seiner et al. [15] have given a streamwise vorticity, and a surface ramp has been installed down- 123
58 detailed review on the mixing enhancement devices in scramjet stream a sonic transverse jet to reduce the low-pressure region 124
59 engines. behind the jet [29]. Nowadays, the transverse injection scheme 125
60 has been utilized in the thermal protection system of the hyper- 126
61 sonic vehicle [30], see Fig. 3, as well as its typical application in 127
62 * Tel.: +86 731 84576447; fax: +8 731 84512301. the attitude control of hypersonic missiles [31–34], and it is able 128
63 E-mail address: gladrain2001@163.com. to recast the bow shock wave into a conical shock wave without 129
64 130
http://dx.doi.org/10.1016/j.ast.2016.01.001
65 131
1270-9638/© 2016 Published by Elsevier Masson SAS.
66 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.2 (1-13)
2 W. Huang / Aerospace Science and Technology ••• (••••) •••–•••

1 67
2 68
3 69
4 70
5 71
6 72
7 73
8 74
9 75
10 76
11 77
12 78
13 79
14 80
15 Fig. 1. Composite figure of the combustion flow field in the HyShot II combustor, (a) wall pressure and an iso-surface of the H2 mass fraction, (b) axial velocity cut through a 81
16 fuel injector, (c) iso-surface of the second invariant of the velocity gradient, λ2 , colored by the temperature and (d) iso-surfaces of the H2 mass fraction (gray) and the heat 82
17 release conditioned on λ2 [17]. 83
18 84
19 85
20 86
21 87
22 88
23 89
24 90
25 91
26 92
27 93
28 94
29 95
30 96
31 97
32 98
33 99
34 100
35 101
36 102
Fig. 2. Schematic of the three-dimensional unsteady vertical structures formed in the HyShot II combustor [21].
37 103
38 104
39 any shock wave/shock wave interaction occurring at the should of 105
40 blunt bodies. In 2015, Huang [35] has provided a detailed review 106
41 on this topic, and the drag and heat release reduction induced by 107
42 a counterflowing jet and its combinations has been summarized. 108
43 The combinatorial configurations include the combination of the 109
44 counterflowing jet and a forward-facing cavity, the combination 110
45 of the counterflowing jet and an aerospike, and the combination 111
46 of the counterflowing jet and energy deposition. Further, he and 112
47 his coworkers have investigated the drag reduction mechanism in- 113
48 duced by a combinational opposing jet and spike concept [36], as 114
49 well as the drag and heat reduction mechanism induced by the 115
50 combinational opposing jet and acoustic cavity concept [37]. 116
51 The transverse injection can provide rapider near-field mixing 117
52 and better fuel penetration capability, and the recirculation region 118
53 induced by the injection can hold the flame. Additionally, it does 119
54 not need more cooling and cannot generate more drag force. How- 120
55 ever, it would generate complex flow field structure and strong 121
56 shock wave, as well as large total pressure loss. The total pres- 122
57 sure loss is not preferable because it leads to the thrust loss, and 123
Fig. 3. Operational principals of Non-ablative Thermal Protection System (NaTPS) for
58 it grows with the increase of the injection angle [38]. At the same aerodynamic force and heat reduction [30].
124
59 time, the compression effect is not beneficial to the vortex gen- 125
60 eration and shedding on the mixing layer between the fuel and 126
tor number and the injection angle, and the multiobjective design
61 air, and this restricts the entrainment mixing and slows down the 127
optimization approach has been proposed to be applied in the de-
62 far-field mixing and combustion process. Therefore, the supersonic 128
63 mixing with very rapid mixing and lower total pressure loss ra- sign process of the transverse injection strategy for the first time. 129
64 tio is highly requested [38]. Huang and Yan [39] have provided Further, they have obtained the Pareto fronts for the optimization 130
65 a survey on the transverse injection from four aspects, namely the of the two- and three-dimensional transverse injection flow fields, 131
66 jet-to-crossflow pressure ratio, the injector configuration, the injec- see Fig. 4, as well as that for a cantilevered ramp injector flow field 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.3 (1-13)
W. Huang / Aerospace Science and Technology ••• (••••) •••–••• 3

1 67
2 68
3 69
4 70
5 71
6 72
7 73
8 74
9 75
10 76
11 77
12 78
13 79
14 80
15 81
16 82
17 83
18 84
19 85
Fig. 4. Pareto fronts for the optimization of the two- and three-dimensional transverse injection flow fields.
20 86
21 87
22 88
23 89
24 90
25 91
26 92
27 93
28 94
29 95
30 96
31 97
32 Fig. 5. Time (200 μs) and line-of-sight integrated OH∗ chemiluminescence imaging comparison of the overall flame structure [54]. 98
33 99
34 [40], and the objective functions should be compromised in order 2. Basic principle 100
35 to obtain the optimized configuration [41,42]. However, in this re- 101
36 view, they have not summarized the basic mixing augmentation They are many influencing factors on the mixing and combus- 102
37 mechanism induced by the combination of the transverse injec- tion process in the transverse injection flow field, i.e. the injector 103
38 tion and other techniques, as well as its research progress, and it configuration [45–48], the injection angle [49,50], the number of 104
39 is a hot topic for the flowpath design of the scramjet engine. At injectors, the injectant species [51], the distance between injec- 105
40 the same time, Mahesh [43] has reviewed the fundamental un- tors [52], and the jet-to-crossflow momentum flux ratio which is 106
41 derstanding of the physical behavior of single-phase, nonreacting defined as follows [53]. The counter–rotating vortex pair (CVP) is 107
42 transverse jets in incompressible and compressible regimes, and the dominant mean flow feature of the jet plume downstream of 108
43 Hassan et al. [44] have reviewed and compared the turbulence the injection location, and the larger value of jet-to-crossflow mo- 109
44 modeling techniques in terms of their performance in predicting mentum flux ratio is beneficial to the evolution of the CVP, as well 110
45 results consistent with the experimental data in the supersonic jet- as the characteristics of the ignition and flame stabilization [54], 111
46 in-crossflow problem, i.e. RANS, LES, and hybrid RANS/LES, and the see Fig. 5. A lateral counter–rotating vortex pair (LCVP) is formed 112
47 hybrid RANS/LES approach is proved to be an improvement over a by controlling the injector pressure in the diamond-shaped port 113
48 pure RANS approach. Meanwhile, its computational time is saved flow field, and it spans the width of the barrel shock wave for gas- 114
49 when compared with the LES approach. dynamically induced flame holding [55–57]. The jet plume from 115
50 Recently, in order to improve the penetration depth and mixing the diamond-shaped injector with high pressure ratios expands 116
51 efficiency of the transverse injection strategy, as well as the total mainly in the spanwise direction [58], and the near-field flow for 117
52 pressure loss and local heat load reduction, many researches have gaseous injection through diamond shaped ports has been stud- 118
53 contributed to the multiport injection and the interaction between ied experimentally by Bowersox et al. [59]. The ability to attain 119
54 jet and vortex generator. At the same time, the high tempera- the pressure rise in near-field for the diamond-shaped orifice is 120
55 ture and high pressure condition in the reacting flow field would weaker than the circular orifice, but it can be improved by instal- 121
56 make a great difference to the fuel injection environment, and lation of a cavity flameholder and introduction of cavity fueling 122
57 the shock wave train and the separation flow in the combustor [60]. One drawback of the transverse injection is the total pressure 123
58 would interact with the fuel injection. Then, the mixing charac- loss due to the relatively strong bow shock wave generated just 124
59 teristics of the transverse injection would vary accordingly when ahead of the injector [61]. 125
60 compared with those obtained under the uniform boundary con- 2 126
(γ P M ) j
61 dition. The purpose of this work is to review the latest advances J = ρ j U 2j /ρ∞ U ∞
2
= (1) 127
62 in this area, consequently, to facilitate the development of novel (γ P M 2 )∞ 128
63 mixing enhancement devices in airbreathing hypersonic propulsion Herein, the subscripts j and ∞ represent the jet exit and the cross- 129
64 systems. The number of works devoted to transverse injection is flow conditions, respectively. ρ and U are density and velocity 130
65 quite large. Therefore, this survey cannot attempt to cite all avail- respectively. γ is the specific heat ratio, P is the pressure and M 131
66 able papers. is the Mach number. 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.4 (1-13)
4 W. Huang / Aerospace Science and Technology ••• (••••) •••–•••

1 tion processes [68], and the “visually observable” jet upper edge in 67
2 the numerical schlieren pattern is usually utilized to describe the 68
3 jet penetration properties, i.e. ∇ ρ  = 0.5 proposed in Ref. [70]. 69
4 The jet penetration for various conditions has been scaled by J D 70
5 [71] and by J 1/2 D [69] for governing length scales in the past, and 71
6 this may imply that the injector diameter D is an important fac- 72
7 tor controlling the jet penetration in the averaged field as well [3]. 73
8 However, the jet-to-crossflow momentum flux ratio has no notable 74
9 impact on the features of turbulent structure [3], not as the injec- 75
10 tant species [72] and the crossflow conditions, and the jet mixing 76
11 is dominated by highly turbulent motions and the magnitude of 77
12 the convection velocity, which is consistent with the frequent ap- 78
13 pearance of the large scale eddies. The convection Mach numbers 79
14 are defined as follows [73]: 80
15 81
16 Fig. 6. Comparison of temperature gradient contours. Top: steady-state RANS result. U1 − Uc Uc − U2 82
Bottom: unsteady result [74]. M c1 = ; M c2 = (3)
17 a1 a2 83
18 84
In the transverse injection flow field, the surface pressure pro- Herein, U 1 and U 2 are the velocity of high- and low-speed streams
19 85
files, the boundary layer separation location, the jet plume height respectively, and U c is a convective speed of the large scale struc-
20 86
and descriptions of recirculation zones and flow structure up- ture. a1 and a2 are the speed of sound of high- and low-speed
21 87
stream and downstream of the jet have been employed to validate streams respectively.
22 88
the predicted results obtained by the computational fluid dynam- A mean RANS profile without velocity fluctuations does not re-
23 89
ics (CFD) approaches [62]. A new penetration correlation has been sult in the proper turbulence levels or mixing in the jet plume,
24 90
proposed by Portz and Segal [63], and it incorporates air Mach especially for the low-angled injection case, see Fig. 6, and the
25 91
number dependence and removes the differences between several most intensively fluctuating region appears on the 50%-averaged-
26 92
existing formulas. Cassell [64] has provided a theoretical estimate concentration track irrespective of the injection conditions.
27 93
28 for the Mach disc height, and it is shown in the following equation. 94
29  
3. Single injection
95
2p 0jet γjet (γjet +1/γjet −1)
30 htheo 1 2 2 96
= The single injection scheme is the simplest strategy for the
31 djet M∞ C d p ∞ γ∞ γjet − 1 γjet + 1 97
mixing process in the scramjet engine, and it has been studied
32  (γjet −1/γjet )
1/4 98
33 p∞ theoretically and experimentally in order to explore its physi- 99
× 1− (2) cal complexity. Lazar et al. [24] have analyzed the influence of
34 p 0jet 100
35
the energy pulse from a Q-switched neodymium-doped–yttrium- 101
36
Herein, C d is the non-dimensional discharge coefficient of the sonic aluminum-garnet laser on the formation of vortices, as well as the 102
37
jet with values around 0.96–0.98. Lin et al. [65] have compared the flow field perturbation. Due to the blockage of the fuel injection, 103
38
measured quantitative ethylene concentration contours within the a bundle of bow shock waves is generated just ahead of the in- 104
39
fuel plumes obtained by Raman scattering with the predicted re- jector. The position of the bow shock waves has been influenced 105
40
sults from the revised jet penetration analysis (JETPEN) code, and by the large scale vortices formed at the interface between the 106
41
this code was first developed by Billig et al. [66] in 1971. Malmuth freestream and jet [75]. Simultaneously, the separation and back- 107
42 and Fedorov [67] have utilized approximations resembling thin wash take place in the boundary layer [76], see Fig. 7, and this is 108
43 shock-layer theory in the Newtonian limit to analyze the trans- propitious to combust. Fig. 7 depicts the schematic diagram of the 109
44 verse fuel injection into a supersonic crossflow. transverse jet in supersonic crossflows, and the sonic jet expands 110
45 The jet’s penetration into the supersonic crossflow is principally through a Prandtl–Meyer fan and then is compressed by the bar- 111
46 dependent on the jet-to-crossflow momentum flux ratio when the rel shock wave and the Mach disk. Downstream of the jet, the flow 112
47 injector diameter is fixed [69], as well as many of the combus- is turned parallel to the wall surface, and a recompression shock 113
48 114
49 115
50 116
51 117
52 118
53 119
54 120
55 121
56 122
57 123
58 124
59 125
60 126
61 127
62 128
63 129
64 130
65 131
66 Fig. 7. Schematic diagram of the transverse jet in supersonic crossflows [77]. 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.5 (1-13)
W. Huang / Aerospace Science and Technology ••• (••••) •••–••• 5

1 67
2 68
3 69
4 70
5 71
6 72
7 73
8 74
9 75
10 76
11 77
12 78
13 79
Fig. 8. Jet penetration profiles: a) circular, 90◦ injector, He, (b) circular, 15◦ injector, He and (c) elliptical, 90◦ injector, He [61].
14 80
15 81
Table 1
16 82
Information for the transverse injection flow fields investigated in partial literature.
17 83
Authors Mach Dimension Injector Injection angle Injectant Jet-to-freestream Category
18 84
number configurations momentum flux
19 ratio 85
20 You et al. [21,22] 2.4 Three Circular 90◦ H2 0.35 DES 86
21 Watanabe et al. [25] 1.9 Three Circular 90◦ Air 1.9 LES 87
22 Hariharan and Babu [47] 3.0 Three Circular, wedge, 90◦ He 0.4165, 0.2466, FANS 88
diamond, chevron 0.4151, 0.4153
23 Srinivasan and Bowersox [56] 2.0, 5.0 Three Diamond 45◦ , 90◦ , 27.5–90◦ , Air 0.4, 2.1, 2.7 GASP flow solver 89
24 45–90◦ 90
25 Gruber et al. [61] 2.0 Three Circular, elliptical 15◦ , 90◦ CO2 , He 1, 2, 3 Experiment 91
Lin et al. [65] 2.0 Three Circular 30◦ , 90◦ C2 H4 0.25 ∼ 6.0 Experiment
26 Watanabe et al. [72] 1.9 Three Circular 90◦ H2 , He, N2 , C2 H4 1.9 LES 92
27 Peterson et al. [74] 2.0 Three Circular 30◦ , 90◦ C2 H4 0.5, 1.0, 1.5 Hybrid RANS/LES 93
28 Gruber et al. [75] 1.98 Three Circular, elliptical 90◦ Air 2.90 Experiment 94
Yang et al. [76] 2.0 Three Circular 30◦ , 45◦ , 60◦ , Aviation kerosene 1a , 2.146a , 3.293a Experiment
29 95
120◦ , 135◦ , 150◦ RP-3
30 Chenault et al. [78] 3.0 Three Circular 25◦ Air 5.12 ISAAC (FANS) 96
31 Sakima et al. [82] 1.81 Three Wedge, circular 90◦ H2 1.0 Experiment 97
Wang et al. [83] 1.6 Three Circular, elliptical 90◦ H2 1.7 LES
32 Kawai et al. [88] 1.6 Three Circular 90◦ Air 1.7 LES 98
33 Peterson et al. [89] 1.6 Three Circular 90◦ Air 1.7 IDDES vs. DES97 99
34 Sriram and Mathew [93] 1.60, 3.75 Three Circular 90◦ Air, N2 1.73, 1.20, 1.70, 2.21, FANS 100
0.61, 1.26, 1.87
35 101
Schetz et al. [122] 4 Three Circular 30◦ He, CH4 , Air 2.1 Experiment
36 Won et al. [124] 3.38 Three Circular 90◦ H2 1.4 DES 102
37 Note: ‘a’ denotes the driven pressure ratio. 103
38 104
39 105
wave is produced, as well as a corresponding recirculation zone. pect ratio in low speed flows for the elliptical injection [80,81].
40 106
The recompression shock-induced vortices are generated through A catalytic reaction on a platinum wire has been utilized to eval-
41 107
the combined influences of the inflow air upwash downstream uate a spatial mixing condition, and the jet plume in the cases of
42 108
of the plume and the mirrored oblique barrel shock wave [78]. wedge injectors penetrates higher than that of the circular injector
43 109
That is to say, four recirculation zones are usually found, termed case [82].
44 110
primary upstream vortices, secondary upstream vortices, primary At the same time, the region of the horseshoe vortex in the el-
45 111
downstream vortices and secondary downstream vortices [79]. liptical injection, corresponding to the recirculation zone detected
46 112
The transverse injection flow field has been investigated from upstream of the injection port, is proved to be somewhat smaller
47 113
several points of view with instantaneous, mean, standard devi- than that in the circular injection [83], and the horseshoe vortex is
48 114
ation (STD), and intensity probability density functions. The Mie related to the flame-holding capability and the hot-spot phenom-
49 scattering technique, which uses the ice crystal with its diame- ena in the combustion situation [84,75]. However, the secondary 115
50 ter being 0.02 μm for scattering, has been utilized by Gruber et vortex between the horseshoe vortex and the jet, which is called 116
51 al. [61] to analyze the influence of the jet-to-crossflow momen- the “hovering vortex” in the low speed flow [85], disappears in the 117
52 tum flux ratio, the injector geometry and the injection species on elliptical injection, see Fig. 9. 118
53 the global flowfield characteristics, and they have found that in- Computational fluid dynamics plays an important role in the 119
54 creases in jet-to-crossflow momentum flux ratio would result in design and assessment of the fuel injection strategies, and this is 120
55 increased jet penetration into the freestream irrespective of the in- due to the high costs involved in flight testing and ground ex- 121
56 jector geometry and the injection angle, see Fig. 8. Fig. 8 depicts perimental tests, as well as the difficulty with making measure- 122
57 the jet penetration profiles for the He injection cases with different ments. Reynolds-averaged Navier–Stokes (RANS) simulations have 123
58 injector geometries and injection angles, and the jet penetration been employed to capture many of the mean flow features [86], 124
59 profiles for the circular cases are nearly the same as those for the and the large-eddy simulation (LES) [87,88] has utilized to include 125
60 elliptical cases. The circular jet has been the subject of many ex- the unsteady structure of the jet plume, as well as the generation 126
61 perimental and computational studies found in the literature, see of mixing and combustion processes [89]. The data set obtained 127
62 Table 1, and Table 1 presents the information of the transverse by Spaid and Zukoski [90] is more suitable for turbulence model 128
63 injection flow fields investigated in partial literature. However, the validation than that obtained by Aso et al. [91], and it has been 129
64 jet from the elliptical injector spreads more rapidly in the spanwise employed as a basis by Huang et al. [92], Sriram and Mathew [79], 130
65 direction, and the injectors with low aspect ratio can augment the Chenault and Beran [62] and Sriram and Mathew [93] to validate 131
66 mixing rates and jet penetrations in regard to those with high as- the influence of turbulence model on the simulation of the two- 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.6 (1-13)
6 W. Huang / Aerospace Science and Technology ••• (••••) •••–•••

1 67
2 68
3 69
4 70
5 71
6 72
7 73
8 74
9 75
10 76
11 77
12 78
13 79
14 80
15 81
16 82
Fig. 9. Mean numerical schlieren-like visualization by contours of ∇ ρ  and streamlines of the mean flow in the y / D = 0 plane for (a) circular and (b) elliptical injections [83].
17 83
18 84
dimensional transverse injection flow field properties. The Aso et
19 85
al.’s configuration owns three-dimensional effects in the flow field
20 86
and strong leading-edge shock waves, and Rizzetta [94] has uti-
21 87
lized it to assess the effect of the compressibility correction, see
22 88
Fig. 10. The predicted results obtained with the compressibility
23 89
correction show better agreement with the experimental data, and
24 90
the peak pressure level upstream of the jet has been reduced. The
25 91
compressibility effect seems to weaken in the far field. Further,
26 92
the Schmidt number is proved to be an important impact on the
27 93
simulation results [95–97], as well as the turbulence model [98],
28 94
and a turbulent Schmidt number of 0.2 is recommended for jet-in-
29 95
crossflow simulations. Fig. 10. Effect of compressibility correction on surface static pressure distribution
30 for w = 1.0 mm and P j / P 0 = 0.310 [94]. 96
The single-time two-point spatial correlation analysis has been
31 97
widely used to clarify the behavior of the turbulent structure [99,
32 98
100], and the correlation coefficient r ( y ,  z) based on the con-
33 99
centration fluctuation is calculated from the following equation:
34 100
35 101
1 
N
C ( y , z) = C i ( y , z), C i ( y , z) = C i ( y , z) − C ( y , z),
36 102
37 N 103
i =1
38  104

39 1 
N 105
40 C rms ( y , z) =


C i 2 ( y , z), i = 1, 2, . . . , N 106
41 N 107
i =1
42 N 
108
1
N i =1 [ C i ( y , z ) · C i ( y +  y , z +  z)]
r ( y ,  z) =
43 109
 ( y , z) · C  ( y +  y , z +  z) (4)
44 C rms rms 110
45 111
46
Herein, point ( y , z) is the reference used for features correlation, 112
47
C i ( y , z) is the instantaneous concentration, C i ( y , z) is its fluctu- 113
ation, C ( y , z)  ( y , z) is the
is the average of concentration, C rms
48 114
49 standard deviation of concentration fluctuation, and  y and  z 115
50 are the spatial differences in y and z directions, respectively. 116
51 117
Fig. 11. Global normalized drag vs. injectant mass flow [109].
52 4. Multiport injection 118
53 119
combustors [106]. At the same time, it would enhance wall cool-
54 The mixing and combustion properties of a multiport injection 120
ing along with reducing total pressure losses and injection-induced
55 scheme are very different from those of a single injection scheme, 121
shock generation [107]. Further, a smaller secondary jet has been
56 and the blockage effects because of the momentum flux of the 122
employed downstream of the primary jet by Viti et al. [108] to
57 front injection flow have a great impact on the rear injection flow 123
58 [101], as well as the preheating effects due to the chemical re- increase the normal force on the flat plate and decrease the nose- 124
59 actions of the front injection flow on the combustion process of down pitching moment. 125
60 the rear injection flow [102]. The accurate fuel distributions in the Pudsey et al. [109] have investigated the film-cooling drag 126
61 flowpath are very crucial for the design of the injection system reduction performance of a small-scale multiport injector array 127
62 [103,104], and a staged-injection configuration can enhance mix- numerically, and the interactions between the adjacent jets are 128
63 ing of primary fuel jet by the interaction between two adjacent responsible for variations in downstream performance, i.e. drag 129
64 jets as well as to promote penetration induced by the secondary reduction, heat transfer, mixing efficiency. There are large three- 130
65 jet’s push up to the primary jet [105]. Thus, it has been combined dimensional recirculation zones exist in this region, and the ef- 131
66 with any other flame holder in the flowpath design of scramjet fective upstream flow for each injector would influence the gen- 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.7 (1-13)
W. Huang / Aerospace Science and Technology ••• (••••) •••–••• 7

1 67
2 68
3 69
4 70
5 71
6 72
7 73
8 74
9 75
10 76
11 77
12 78
13 79
14 80
15 81
16 82
17 Fig. 12. Schematic diagram of multiport injection flow field. 83
18 84
19 85
20 86
21 87
22 88
23 89
24 90
25 91
26 92
27 93
28 94
29 95
30 96
31 97
32 98
Fig. 13. Comparison of maximum trajectories between single and staged injections
33 99
[105].
34 100
35 101
36
eration of normal CVPs. They have found that the global drag 102
37
reduction has a strong linear dependence on the injection mass Fig. 14. Penetration height comparison at x/ D = 10 [105]. 103
38
flow rate, see Fig. 11, and the drag decreases with the increase of 104
39
the flow rate, as well as the heat transfer rate and the mixing effi- tion species and secondary pulsed-injection. They have found that 105
40 ciency. However, it has only a slight dependence on the streamwise the air–air case shows higher penetration compared with others 106
41 spacing, and it generally increases with the increase of streamwise after the secondary injection, see Fig. 13, and the effect of sec- 107
42 spacing, as well as the heat transfer and the mixing efficiency. Re- ondary injection occurs earlier in the air-primary injection. At the 108
43 cently, they have investigated the flow field properties of boundary same time, there is a hysteresis exists in penetration height for 109
44 layer combustion using a multiporthole injector array numerically the pulsed injection, see Fig. 14, and the ascending phase is ben- 110
45 [110], and the influences of the injectant mass flow rate and the eficial for the fuel penetration. This phenomenon is the same as 111
46 streamwise jet-to-jet spacing have been evaluated [110,111]. They that observed in the mode transition process [112]. Davitian et al. 112
47 found that the ignition length increases with the decrease of the [113] have explored the stability properties of the shear layer for 113
48 mass flow rate when the injection mass flow rate is high, and the the single gaseous jet in crossflow or transverse jet in order to 114
49 wall heat transfer rate reduction induced by the boundary layer control the jet penetration and spread, and the periodic jet forc- 115
50 combustion is beneficial for the flowpath design of scramjet en- ing has been proposed as well. The controlled supersonic swirling 116
51 gine. injector (CSSI) has been proved to be able to achieve good mixing 117
52 Lee [101,102] has investigated the mixing and combustion char- and entrainment as well [114,115]. However, the staged-injection 118
53 acteristics of dual transverse injection in scramjet combustor com- does not own significant advantage in mixing, and this conclusion 119
54 prehensively, see Fig. 12, and the influences of the jet-to-crossflow is different from that obtained by Gao and Lee [26]. This may be 120
55 momentum flux ratio and the injector distance have been consid- induced by the differences in the geometric configuration and the 121
56 ered. He has found that the dual injection system is beneficial to boundary conditions employed. 122
57 the improvement of the mixing and combustion efficiency, as well 123
58 as the penetration depth, and it would induce much more stagna- 124
5. Interaction between jet and vortex generator
59 tion pressure losses. At the same time, there exist an optimal in- 125
60 jector distance for the improvement of the mixing and combustion 126
61 efficiency, and it increases with the increase of the jet-to-crossflow The vortex generator has been employed to promote the pen- 127
62 momentum flux ratio. etration of the fuel into the core airstream, i.e. the pylon and the 128
63 Takahashi et al. [105] have extended the fluorescence ratio hypermixer, and it would not bring high pressure losses, as well 129
64 technique for processing planar laser-induced fluorescence data for as the occurrence of the flashback phenomenon. Vinogradov et al. 130
65 quantitative imaging of the injectant mole-fraction and density in [116] have provided a comprehensive review on the thin pylons 131
66 staged injection configurations, as well as the influences of injec- applications. The low-pressure region behind the pylon leads to in- 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.8 (1-13)
8 W. Huang / Aerospace Science and Technology ••• (••••) •••–•••

1 creased penetration, and the weak vertical structures induced help Kim et al. [120] have explored the relationship between a wall- 67
2 enhance spreading and mixing. mounted alternating-ramp-wedge-type hypermixer and transverse 68
3 Wang et al. have investigated the vertical structures over a injection compared with a step mixer experimentally, see Fig. 16, 69
4 delta-wing [117], a hemisphere [118] and a finite cylinder [119] and the mixing performance (i.e. penetration height, vorticity, and 70
5 with nanoparticle-based planar laser scattering (NPLS), and the turbulent stress) is improved when a transverse jet is injected in 71
6 characteristics of periodicity and similar geometry of the spanwise the vertical region. In their research work, the step mixer has been 72
7 and streamwise large scale vortex structures have been obtained. utilized only as the baseline model. 73
8 74
In the flow structures over a delta-wing, the primary vortex pair
9
would induce the low speed flow in the boundary layer and gen- 6. Interaction between jet and shock wave 75
10 76
erate the secondary instability vortex pair, see Fig. 15. The regions
11
just downstream of these configurations all provide a suitable The OpenFOAM CFD toolbox has been employed by Shekarian 77
12
place for the fuel injection, although these configurations would et al. [121] to investigate the influence of incident shock wave 78
13
induce additional total pressure loss. However, the interaction be- and its location on mixing and flame holding of hydrogen in su- 79
14 personic airstream, and the angle of the wedge which is used to 80
tween the fuel injection and these vortex generators has rarely
15 generate the incident shock wave keeps constant. The interaction 81
been investigated in the open literature to the best of the author’s
16 between the mixing layer and an oblique shock wave generates 82
knowledge, and it should be explored in the near future.
17 strong axial vortices that stretch the fuel/air interface, and it could 83
18 reduce penetration and increase mixing for injectants of all molec- 84
19 ular weights [122]. The recirculation region downstream of the 85
20 injector is expanded downstream when the incident shock wave 86
21 hits at downstream of the injector, and this increases the residence 87
22 time at this location and consequently enhances the mixing be- 88
23 tween the fuel and supersonic airstream, as well as the decreasing 89
24 of the fuel concentration in the positive direction of x-axis [121]. 90
25 This implies that the incident shock wave has a great impact on 91
26 the hydrogen concentration reduction, and the recirculation region 92
27 downstream of the injector has an important impact on the mixing 93
28 and flame holding capability of the fuel. 94
29 The OH mass fraction is the useful flame marker to study the 95
30 influence of incident shock wave on flame holding capability [123], 96
31 as well as the locations of self-ignition [124], see Fig. 17, and 97
32 Fig. 17 represents the comparison of OH mass fraction contours 98
33 for the cases without and with incident shock wave. OH radical is 99
34 developed downstream with the incident shock wave introduced 100
35 downstream of the injector, however, it exists only in a limited 101
36 area without the incident shock wave. This implies that the inci- 102
37 dent shock wave is beneficial to the flame stability, and it could 103
38 Fig. 15. Crosswise NPLS images in the y–z (x = 88 mm) plane for the delta-wing on work as flame-holder, see Fig. 18. In Fig. 18, it is clearly observed 104
39 a flat plate [117]. that the shock wave impingement could enhance the shock in- 105
40 106
41 107
42 108
43 109
44 110
45 111
46 112
47 113
48 114
49 115
50 116
51 117
52 118
53 119
54 120
Fig. 16. Configuration of the two mixer models [120].
55 121
56 122
57 123
58 124
59 125
60 126
61 127
62 128
63 129
64 130
65 131
66 Fig. 17. Comparison of OH mass fraction contours [121]. 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.9 (1-13)
W. Huang / Aerospace Science and Technology ••• (••••) •••–••• 9

1 time, they only considered the two-dimensional flow field and 67


2 the simple H2 –O2 chemical reaction mechanism, and more work 68
3 should be conducted in the near future. 69
4 Erdem et al. [126] have used the Particle Image Velocimetry 70
5 (PIV) to analyze the influence of an impinging shock wave on the 71
6 jet mixing with the jet-to-mainstream momentum flux ratio be- 72
7 ing 2.75, see Fig. 19, and the extent of jet spreading is enhanced 73
8 in the near field when the shock wave impingement location is 74
9 shifted downstream. At the same time, the downstream recom- 75
10 pression shock wave becomes apparent as well. 76
11 Recently, Huang et al. [127] have investigated the interaction 77
12 between the incident shock wave and the transverse slot injec- 78
13 tion flow field with low and high jet-to-crossflow pressure ratios 79
14 numerically, and the transverse slot injection flow field is dis- 80
Fig. 18. Combined image of ultraviolet and schlieren image in the shock wave im-
15
pingement flow field [123].
turbed seriously with the increase of the swept angle of the ramp. 81
16 Meanwhile, the separation zones upstream and downstream of the 82
17 duced separated region. Recently, Tahsini and Mousavi [125] have injection port increase with the increase of the intensity of the in- 83
18 studied the effect of oblique shock wave on combustion efficiency cident shock wave, and the boundary layer properties vary sharply. 84
19 of hydrogen injection into a supersonic crossflow, and they found Further, they have evaluated the mixing augmentation induced by 85
20 that the shock impinging upstream of the injection slot can ef- the interaction between the shock wave and the transverse jet 86
21 fectively increase the combustion efficiency. However, the pressure [128], and the influences of the jet-to-crossflow pressure ratio and 87
22 loss across the shock wave increases by enhancement of the in- the wedge angle on the mixing efficiency have been investigated, 88
23 duced oblique shock wave strength, and there must be a trade-off see Fig. 20. They found that the mixing efficiency increases with 89
24 between the combustion efficiency and pressure loss. At the same the increase of the wedge angle irrespective of the value of the 90
25 91
26 92
27 93
28 94
29 95
30 96
31 97
32 98
33 99
34 100
35 101
36 102
37 103
38 104
39 105
40 106
41 107
42 108
43 109
44 110
45 111
46 112
47 113
48 Fig. 19. Comparison of averaged flow field turbulence intensity contours [126]. 114
49 115
50 116
51 117
52 118
53 119
54 120
55 121
56 122
57 123
58 124
59 125
60 126
61 127
62 128
63 129
64 130
65 Fig. 20. Mixing efficiency comparisons for the cases with different wedge angles and different jet-to-crossflow pressure ratios, (a) P j / P ∞ = 10.29, (b) P j / P ∞ = 17.72 and 131
66 (c) P j / P ∞ = 25.15 [128]. 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.10 (1-13)
10 W. Huang / Aerospace Science and Technology ••• (••••) •••–•••

1 67
2 68
3 69
4 70
5 71
6 72
7 73
8 74
9 75
10 76
11 77
12 78
13 79
14 80
15 81
16 82
17 83
18 84
Fig. 21. Schematic diagram of the flow field properties around the transverse micro-jet under influence of the oblique shock wave [129].
19 85
20 86
21
sting, and there are some novel vortex structures exist for en- 87
22
hancement of the mixing and combustion performance. At the 88
23 same time, the numerical and experimental study on the in- 89
24 fluence of the operating and structural parameter variations 90
25 is the most popular for the researchers at present, and the 91
26 refined vortex structure capture has attracted an increasing 92
27 attention. 93
28 • The flowpath design of the scramjet engine is very crucial for 94
29 the improvement of its overall performance, and thus, the ac- 95
Fig. 22. Schlieren photograph of interaction between a pseudo shock wave and fuel
30 injection [132]. curate fuel distributions become important. That is to say, the 96
31 multiport injection scheme is commonly encountered in the 97
32 jet-to-crossflow pressure ratio. At the same time, Gerdroodbary scramjet engine, and its arrangement should be optimized in 98
33 and his coworkers [129] have investigated the influence of shock order to obtain the relative better performance, as well as the 99
34 wave position on sonic transverse hydrogen micro-jets in super- interaction mechanism between the adjacent injections and 100
35 sonic crossflow numerically, see Fig. 21, and both the penetration the interaction between the incident shock wave and the fuel 101
36 and mixing hydrogen for different shock wave positions under a injection. 102
37 specified set of flow conditions have been evaluated, as well as • The recirculation region formed just downstream of the vor- 103
38 any relationships between mass distribution and different configu- tex generators is very promising for the enhancement of the 104
39 rations. mixing and combustion performance, although the vortex gen- 105
40 The stereoscopic particle image velocimetry (PIV) technique has erators would induce additional total pressure loss, as well 106
41 been used to study the interaction between a pseudo shock wave as more drag force. The combinatorial performance between 107
42 (PSW) observed commonly in the isolator [130] and fuel mix- the transverse injection and the vortex generators should be 108
43 ing enhancement [131], see Fig. 22, and the complicated three- optimized, as well as the capture of the interaction mecha- 109
44 dimensional distorting motions induced by the PSW strongly dis- nism. 110
45 turb the transverse injection flow field and improve the mixing 111
46 between the supersonic freestream and fuel. In 2015, Desikan Conflict of interest statement 112
47 et al. [133] have investigated the interaction of a highly under- 113
48 expanded supersonic jet with hypersonic cross flow experimen- None declared. 114
49 tally and numerically, and the wall static pressure measurement, 115
50 Schlieren, and oil flow visualization approaches have been em- Acknowledgements 116
51 ployed in the ground experimental test. These data would provide 117
52 a useful benchmark for this phenomenon. 118
The author would like to express his thanks for the sup-
53 119
port from the National Natural Science Foundation of China
54 7. Remarks and conclusions 120
(No. 11502291) and the Science Foundation of National University
55 121
of Defence Technology (No. JC14-01-01). Also, the author thanks
56 In this survey, the research progress on the transverse jet in 122
Mr. Yan-hui Zhao for the literature provided and the anonymous
57 supersonic crossflows has been summarized systematically from 123
reviewers for some very critical and constructive recommendations
58 four aspects, namely single injection, multiport injection, interac- 124
on this article.
59 tion between jet and vortex generator, and interaction between 125
60 jet and shock wave. At the same time, the basic principle of 126
61 the transverse jet in the supersonic crossflow has been provided References 127
62 in this review as well. We have come to the following conclu- 128
[1] E.T. Curran, Scramjet engines: the first forty years, J. Propuls. Power 17 (6)
63 sions: 129
(2001) 1138–1148.
64 130
[2] W. Huang, M. Pourkashanian, L. Ma, D.B. Ingham, S.B. Luo, Z.G. Wang, Investi-
65 • Many kinds of injector configuration have been proposed and gation on the flameholding mechanisms in supersonic flows: backward-facing
131
66 investigated in the past few years, i.e. diamond, chevron and step and cavity flameholder, J. Vis. 14 (1) (2011) 63–74. 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.11 (1-13)
W. Huang / Aerospace Science and Technology ••• (••••) •••–••• 11

1 [3] J. Watanabe, T. Kouchi, K. Takita, G. Masuya, Characteristics of hydrogen jets [33] J.M. Zhang, Y.D. Cui, J. Cai, H.S. Dou, Numerical investigation of lateral jets 67
2 in supersonic crossflow: large-eddy simulation study, J. Propuls. Power 29 (3) over body of revolution in supersonic crossflow, J. Propuls. Power 28 (1) 68
3 (2013) 661–674. (2012) 33–45. 69
[4] W. Huang, H. Qin, S.B. Luo, Z.G. Wang, Research status of key techniques for [34] H.B. Ebrahimi, Numerical investigation of jet interaction in a supersonic
4 70
shock-induced combustion ramjet (shcramjet) engine, Sci. China, Technol. Sci. freestream, J. Spacecr. Rockets 45 (1) (2008) 95–103.
5 53 (1) (2010) 220–226. [35] W. Huang, A survey of drag and heat reduction in supersonic flows by a 71
6 [5] M.R. Pohlman, R.B. Greendyke, Parametric analysis of pylon-aided fuel injec- counterflowing jet and its combinations, J. Zhejiang Univ. Sci. A 16 (7) (2015) 72
7 tion in scramjet engines, J. Eng. Gas Turbines Power 135 (2013) 024501. 551–561. 73
[6] T.M. Abdel-Salam, T.O. Mohieldin, S.N. Tiwari, Investigation of mixing and flow [36] W. Huang, J. Liu, Z.X. Xia, Drag reduction mechanism induced by a combina-
8 74
characteristics in a dual-mode combustor, in: 41st Aerospace Sciences Meet- tional opposing jet and spike concept in supersonic flows, Acta Astronaut. 115
9 ing and Exhibit, Reno, Nevada, 2003, AIAA Paper 2003-372. 75
(2015) 24–31.
10 [7] B. Wei, B. Chen, M. Yan, X. Shi, Y. Zhang, X. Xu, Operational sensitivities of [37] W. Huang, L. Yan, J. Liu, L. Jin, J.G. Tan, Drag and heat reduction mechanism 76
11 an integrated aerodynamic-ramp-injector/gas-portfire flame holder in a su- in the combinational opposing jet and acoustic cavity concept for hypersonic 77
12 personic combustor, Acta Astronaut. 81 (2012) 102–110. vehicles, Aerosp. Sci. Technol. 42 (2015) 407–414. 78
[8] L. Maddalena, T.L. Campioli, J.A. Schetz, Experimental and computational in- [38] S. Aso, K. Inoue, K. Yamaguchi, Y. Tani, A study on supersonic mixing by
13 79
vestigation of light-gas injectors in Mach 4.0 crossflow, J. Propuls. Power circular nozzle with various injection angles for air breathing engine, Acta
14 22 (5) (2006) 1027–1038. Astronaut. 65 (2009) 687–695. 80
15 [9] P. Manna, R. Behera, D. Chakraborty, Liquid-fueled strut-based scramjet com- [39] W. Huang, L. Yan, Progress in research on mixing techniques for transverse 81
16 bustor design: a computational fluid dynamics approach, J. Propuls. Power injection flow fields in supersonic crossflows, J. Zhejiang Univ. Sci. A 14 (8) 82
24 (2) (2008) 274–281. (2013) 554–564.
17 83
[10] H. Takahashi, Q. Tu, C. Segal, Effects of pylon-aided fuel injection on mixing [40] W. Huang, S.B. Li, L. Yan, J.G. Tan, Multiobjective design optimization of a can-
18 in a supersonic flowfield, J. Propuls. Power 26 (5) (2010) 1092–1101. 84
tilevered ramp injector using the surrogate-assisted evolutionary algorithm,
19 [11] L.D.R. Montes, P.I. King, M.R. Gruber, C.D. Carter, K.Y. Hsu, Mixing effects of J. Aerosp. Eng. 28 (2015) 04014138. 85
20 pylon-aided fuel injection located upstream of a flameholding cavity in su- [41] W. Huang, J. Yang, L. Yan, Multi-objective design optimization of the trans- 86
personic flow, in: 41st AIAA/ASME/SAE/ASEE Joint Propulsion Conference & verse gaseous jet in supersonic flows, Acta Astronaut. 93 (2014) 13–22.
21 87
Exhibit, Tucson, Arizona, 2005, AIAA Paper 2005-3913. [42] W. Huang, Design exploration of three-dimensional transverse jet in a super-
22 [12] F. Vergine, M. Crisanti, L. Maddalena, V. Miller, M. Gamba, Supersonic combus- 88
sonic crossflow based on data mining and multi-objective design optimization
23 tion of pylon-injected hydrogen in high-enthalpy flow with imposed vortex 89
approaches, Int. J. Hydrog. Energy 39 (2014) 3914–3925.
24 dynamics, J. Propuls. Power 31 (1) (2015) 89–103. [43] K. Mahesh, The interaction of jets with crossflow, Annu. Rev. Fluid Mech. 45 90
25 [13] W. Huang, S.B. Li, L. Yan, Z.G. Wang, Performance evaluation and parametric (2013) 379–407. 91
analysis on cantilevered ramp injector in supersonic flows, Acta Astronaut. 84 [44] E. Hassan, J. Boles, H. Aono, D. Davis, W. Shyy, Supersonic jet and crossflow
26 92
(2013) 141–152. interaction: computational modeling, Prog. Aerosp. Sci. 57 (2013) 1–24.
27 [14] W. Huang, Z.G. Wang, W.D. Liu, S.B. Li, Influences of angles of attack and 93
[45] T. Kouchi, G. Masuya, K. Hirano, A. Matsuo, S. Tomioka, Supersonic combustion
28 sideslip on the flow field of a cantilevered ramp injector in supersonic flows, 94
using a sting-shaped fuel injector, J. Propuls. Power 29 (3) (2013) 639–647.
29 Trans. Jpn. Soc. Aeronaut. Space Sci. 57 (4) (2014) 195–200. 95
[46] M.R. Gruber, L.P. Goss, Surface pressure measurements in supersonic trans-
[15] J.M. Seiner, S.M. Dash, D.C. Kenzakowski, Historical survey on enhanced mix-
30 verse injection flowfields, J. Propuls. Power 15 (5) (1999) 633–641. 96
ing in scramjet engines, J. Propuls. Power 17 (6) (2001) 1273–1286.
31 [47] A.R. Hariharan, V. Babu, Transverse injection into a supersonic cross flow 97
[16] A. Kovar, E. Schulein, Comparison of experimental and numerical investigation
through a circular injector with chevrons, J. Fluids Eng. 136 (2014) 021204.
32 on a jet in a supersonic cross-flow, Aeronaut. J. (2006) 353–360. 98
[48] M.R. Gruber, A.S. Nejad, T.H. Chen, J.C. Dutton, Transverse injection from cir-
33 [17] C. Fureby, M. Chapuis, E. Fedina, S. Karl, CFD analysis of the HyShot II scramjet 99
cular and elliptic nozzles into a supersonic crossflow, J. Propuls. Power 16 (3)
combustor, Proc. Combust. Inst. 33 (2011) 2399–2405.
34 (2000) 449–457. 100
[18] R. Pecnik, V.E. Terrapon, F. Ham, G. Iaccarino, H. Pitsch, Reynolds-averaged
35 [49] W. Huang, L. Ma, M. Pourkashanian, D.B. Ingham, S.B. Luo, Z.G. Wang, Para- 101
Navier–Stokes simulations of the HyShot II scramjet, AIAA J. 50 (8) (2012)
36 metric effects in a scramjet engine on the interaction between the air stream 102
1717–1732.
and the injection, Proc. Inst. Mech. Eng., G J. Aerosp. Eng. 226 (3) (2012)
37 [19] D. Cecere, A. Ingenito, E. Giacomazzi, L. Romagnosi, C. Bruno, Hydrogen/air 103
294–309.
38 supersonic combustion for future hypersonic vehicles, Int. J. Hydrog. Energy 104
[50] P.S. King, R.H. Thomas, J.A. Schetz, F.S. Billig, Combined tangential-normal in-
36 (2011) 11969–11984.
39 jection into a supersonic flow, J. Propuls. Power 7 (3) (1991) 420–430. 105
[20] Z.A. Rana, B. Thornber, D. Drikakis, Dynamics of sonic hydrogen jet injection
40 [51] W. Huang, J. Liu, L. Jin, L. Yan, Molecular weight and injector configuration 106
and mixing inside scramjet combustor, Eng. Appl. Comput. Fluid Mech. 7 (1)
effects on the transverse injection flow field properties in supersonic flows,
41 (2013) 13–39. 107
Aerosp. Sci. Technol. 32 (2014) 94–102.
42 [21] Y. You, H. Lüdeke, K. Hannemann, On the flow physics of a low momentum 108
[52] W. Huang, Numerical investigation on staged sonic jet interaction mechanism
flux ratio jet in a supersonic turbulent crossflow, Europhys. Lett. 97 (2012)
43 in a supersonic cross flow, Proc. Inst. Mech. Eng., G J. Aerosp. Eng. 228 (14) 109
24001.
44 [22] Y. You, H. Luedeke, K. Hannemann, Injection and mixing in a scramjet com- (2014) 2641–2651. 110
45 bustor: DES and RANS studies, Proc. Combust. Inst. 34 (2) (2012) 2083–2092, [53] A.R. Karagozian, Transverse jets and their control, Prog. Energy Combust. Sci. 111
http://dx.doi.org/10.1016/j.proci.2012.10.001. 36 (2010) 531–553.
46 112
[23] Z.A. Rana, B. Thornber, D. Drikakis, Transverse jet injection into a supersonic [54] M. Gamba, M.G. Mungal, R.K. Hanson, Ignition and near-wall burning in trans-
47 verse hydrogen jets in supersonic crossflow, in: 49th AIAA Aerospace Sciences 113
turbulent cross-flow, Phys. Fluids 23 (2011) 046103.
48 [24] E. Lazar, G. Elliott, N. Glumac, Energy deposition applied to a transverse jet in Meeting including the New Horizons Forum and Aerospace Exposition, Or- 114
49 a supersonic crossflow, AIAA J. 48 (8) (2010) 1662–1672. lando, Florida, 2011, AIAA Paper 2011-319. 115
[25] J. Watanabe, T. Kouchi, K. Takita, G. Masuya, Numerical study on the turbu- [55] K. Kobayashi, R.D.W. Bowersox, R. Srinivasan, N.R. Tichenor, C.D. Carter, M.D.
50 116
lent structure of transverse jet into supersonic flow, AIAA J. 49 (9) (2011) Ryan, Experimental and numerical studies of diamond-shaped injector in a
51 117
2057–2067. supersonic flow, J. Propuls. Power 26 (2) (2010) 373–376.
52 [56] R. Srinivasan, R.D.W. Bowersox, Simulation of transverse gaseous injection 118
[26] Z.X. Gao, C.H. Lee, Numerical research on mixing characteristics of different
53 injection schemes for supersonic transverse jet, Sci. China, Technol. Sci. 54 (4) through diamond ports into supersonic freestream, J. Propuls. Power 23 (4) 119
54 (2011) 883–893. (2007) 772–782. 120
[27] F. Genin, S. Menon, Dynamics of sonic jet injection into supersonic crossflow, [57] R. Srinivasan, R.D.W. Bowersox, Transverse injection through diamond and cir-
55 121
J. Turbul. 11 (4) (2010) 1–30. cular ports into a Mach 5.0 freestream, AIAA J. 46 (8) (2008) 1944–1962.
56 [58] S. Tomioka, L.S. Jacobsen, J.A. Schetz, Sonic injection from diamond-shaped 122
[28] S.H. Lee, T. Mitani, Mixing augmentation of transverse injection in scramjet
57 combustor, J. Propuls. Power 19 (1) (2003) 115–124. orifices into a supersonic crossflow, J. Propuls. Power 19 (1) (2003) 104–114. 123
58 [29] Y.H. Byun, K.J. Bae, S. Wallis, V. Viti, J.A. Schetz, R. Bowersox, Jet interaction in [59] R.D.W. Bowersox, H. Fan, D. Lee, Sonic injection into a Mach 5.0 freestream 124
59 supersonic flow with a downstream surface ramp, J. Spacecr. Rockets 42 (1) through diamond orifices, J. Propuls. Power 20 (2) (2004) 280–287. 125
(2005) 38–44. [60] S. Tomioka, T. Kohchi, R. Masumoto, M. Izumikawa, A. Matsuo, Super-
60 126
[30] Y. Liu, Z. Jiang, Concept of non-ablative thermal protection system for hyper- sonic combustion with supersonic injection through diamond-shaped orifices,
61 sonic vehicles, AIAA J. 51 (3) (2013) 584–590. J. Propuls. Power 27 (6) (2011) 1196–1203. 127
62 [31] B. Stahl, H. Emunds, A. Gülhan, Experimental investigation of hot and cold [61] M.R. Gruber, A.S. Nejad, T.H. Chen, J.C. Dutton, Mixing and penetration studies 128
63 side jet interaction with a supersonic cross-flow, Aerosp. Sci. Technol. 13 of sonic jets in a Mach 2 freestream, J. Propuls. Power 11 (2) (1995) 315–323. 129
(2009) 488–496. [62] C.F. Chenault, P.S. Beran, K –ε Reynolds stress turbulence model comparisons
64 130
[32] B.Y. Min, J.W. Lee, Y.H. Byun, Numerical investigation of the shock interac- for two-dimensional injection flows, AIAA J. 36 (8) (1998) 1401–1412.
65 131
tion effect on the lateral jet controlled missile, Aerosp. Sci. Technol. 10 (2006) [63] R. Portz, C. Segal, Penetration of gaseous jets in supersonic flows, AIAA J.
66 385–393. 44 (10) (2006) 2426–2429. 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.12 (1-13)
12 W. Huang / Aerospace Science and Technology ••• (••••) •••–•••

1 [64] L.A. Cassel, Applying jet interaction technology, J. Spacecr. Rockets 40 (4) [97] J.A. Boles, J.R. Edwards, R.A. Baurle, Large-eddy/Reynolds-averaged Navier– 67
2 (2003) 523–537. Stokes simulations of sonic injection into Mach 2 crossflow, AIAA J. 48 (7) 68
3 [65] K.C. Lin, M. Ryan, C. Carter, M. Gruber, C. Raffoul, Raman scattering measure- (2010) 1444–1456. 69
ments of gaseous ethylene jets in Mach 2 supersonic crossflow, J. Propuls. [98] Z.A. Rana, B. Thornber, D. Drikakis, On the importance of generating accu-
4 70
Power 26 (3) (2010) 503–513. rate turbulent boundary condition for unsteady simulations, J. Turbul. 12 (35)
5 [66] F.S. Billig, R.C. Orth, M. Lasky, A unified analysis of gaseous jet penetration, (2011) 1–39. 71
6 AIAA J. 9 (6) (1971) 1048–1058. [99] K.M. Smith, J.C. Dutton, Investigation of large-scale structures in supersonic 72
7 [67] N. Malmuth, A. Fedorov, Thin shock layer theory model for a jet in a hyper- planar base flows, AIAA J. 34 (6) (1996) 1146–1152. 73
sonic cross flow, in: 43rd AIAA Aerospace Sciences Meeting and Exhibit, Reno, [100] N.L. Messersmith, J.C. Dutton, Characteristic features of large structures in
8 74
Nevada, 2005, AIAA Paper 2005-893. compressible mixing layers, AIAA J. 34 (9) (1996) 1814–1821.
9 [68] M. Gamba, M.G. Mungal, Ignition, flame structure and near-wall burning in [101] S.H. Lee, Characteristics of dual transverse injection in scramjet combustor, 75
10 transverse hydrogen jets in supersonic crossflow, J. Fluid Mech. 780 (2015) part 1: mixing, J. Propuls. Power 22 (5) (2006) 1012–1019. 76
11 226–273. [102] S.H. Lee, Characteristics of dual transverse injection in scramjet combustor, 77
12 [69] D. Papamoschou, D.G. Hubbard, Visual observations of supersonic transverse part 2: combustion, J. Propuls. Power 22 (5) (2006) 1020–1026. 78
jets, Exp. Fluids 14 (1993) 468–476. [103] L.S. Jacobsen, J.A. Schetz, W.F. Ng, Flowfield near a multiport injector array in
13 79
[70] A. Ben-Yakar, R.K. Hanson, Ultra-fast-framing Schlieren system for studies of a supersonic flow, J. Propuls. Power 16 (2) (2000) 216–226.
14 the time evolution of jets in supersonic crossflows, Exp. Fluids 32 (2002) [104] W. Huang, Effect of jet-to-crossflow pressure ratio arrangement on turbulent 80
15 652–666. mixing in a flowpath with square staged injectors, Fuel 144 (2015) 164–170. 81
16 [71] M.R. Gruber, A.S. Nejad, T.H. Chen, J.C. Dutton, Compressibility effects in su- [105] H. Takahashi, S. Ikegami, G. Masuya, M. Hirota, Extended quantitative fluo- 82
personic transverse injection flowfield, Phys. Fluids 9 (5) (1997) 1448–1461. rescence imaging for multicomponent and staged injection into supersonic
17 83
[72] J. Watanabe, T. Kouchi, K. Takita, G. Masuya, Large-eddy simulation of jet in crossflows, J. Propuls. Power 26 (4) (2010) 798–807.
18 supersonic crossflow with different injectant species, AIAA J. 50 (12) (2012) [106] D. Chakraborty, A.P. Roychowdhury, V. Ashok, P. Kumar, Numerical investi- 84
19 2765–2778. gation of staged transverse sonic injection in Mach 2 stream in confined 85
20 [73] H. Takahashi, G. Masuya, M. Hirota, Effects of injection and main flow con- environment, Aeronaut. J. (2003) 719–729. 86
ditions on supersonic turbulent mixing structure, AIAA J. 48 (8) (2010) [107] A.S. Pudsey, R.R. Boyce, Numerical investigation of transverse jets through
21 87
1748–1756. multiport injector arrays in supersonic crossflow, J. Propuls. Power 26 (6)
22 [74] D.M. Peterson, G.V. Candler, Simulations of mixing for normal and low-angled (2010) 1225–1236. 88
23 injection into supersonic crossflow, AIAA J. 49 (12) (2011) 2792–2804. [108] V. Viti, S. Wallis, J.A. Schetz, R. Neel, R.D.W. Bowersox, Jet interaction with a 89
24 [75] M.R. Gruber, A.S. Nejad, T.H. Chen, et al., Bow shock/jet interaction in com- primary jet and an array of smaller jets, AIAA J. 42 (7) (2004) 1358–1368. 90
pressible transverse injection flowfields, AIAA J. 34 (1996) 2191–2193. [109] A.S. Pudsey, R.R. Boyce, V. Wheatley, Hypersonic viscous drag reduction via
25 91
[76] H. Yang, F. Li, B. Sun, Trajectory analysis of fuel injection into supersonic cross multiporthole injector arrays, J. Propuls. Power 29 (5) (2013) 1087–1096.
26 92
flow based on Schlieren method, Chin. J. Aeronaut. 25 (2012) 42–50. [110] A.S. Pudsey, V. Wheatley, R.R. Boyce, Supersonic boundary-layer combustion
27 [77] L. Yan, W. Huang, T. Zhang, H. Li, X.T. Yan, Numerical investigation of the via multiporthole injector arrays, AIAA J. 53 (10) (2015) 2890–2906. 93
28 nonreacting and reacting flow fields in a transverse gaseous injection chan- [111] A.S. Pudsey, V. Wheatley, R.R. Boyce, Behavior of multiple-jet interactions in a 94
29 nel with different species, Acta Astronaut. 17–23 (2014), http://dx.doi.org/ hypersonic boundary layer, J. Propuls. Power 31 (1) (2015) 144–155. 95
10.1016/j.actaastro.2014.08.018. [112] W. Huang, L. Yan, J.G. Tan, Survey on the mode transition technique in com-
30 96
[78] C.F. Chenault, P.S. Beran, R.D.W. Bowersox, Numerical investigation of super- bined cycle propulsion systems, Aerosp. Sci. Technol. 39 (2014) 685–691.
31 sonic injection using a Reynolds-stress turbulence model, AIAA J. 37 (10) [113] J. Davitian, C. Hendrickson, D. Getsinger, R.T. M’Closkey, A.R. Karagozian, 97
32 (1999) 1257–1269. Strategic control of transverse jet shear layer instabilities, AIAA J. 48 (9) 98
33 [79] A.T. Sriram, J. Mathew, Improved prediction of plane transverse jets in super- (2010) 2145–2156. 99
sonic crossflows, AIAA J. 44 (2) (2006) 405–407. [114] S. Murugappan, E. Gutmark, C. Carter, J. Donbar, M. Gruber, K.Y. Hsu, Trans-
34 100
[80] D.S. Liscinsky, B. True, J.D. Holdeman, Crossflow mixing of noncircular jets, verse supersonic controlled swirling jet in a supersonic cross stream, AIAA J.
35 J. Propuls. Power 12 (1996) 225–230. 44 (2) (2006) 290–300. 101
36 [81] T.T. Lim, T.H. New, S.C. Luo, Scaling of trajectories of elliptic jets in crossflow, [115] S. Murugappan, E. Gutmark, C. Carter, Control of penetration and mixing of an 102
37 AIAA J. 44 (2006) 3157–3160. excited supersonic jet into a supersonic cross stream, Phys. Fluids 17 (2005) 103
[82] F. Sakima, T. Arai, J. Kasahara, M. Murakoshi, T. Ami, F. He, H. Sugiyama, Mix- 106101.
38 104
ing of a hydrogen jet from a wedge shaped injector into a supersonic cross [116] V.A. Vinogradov, Y.M. Shikhman, C. Segal, A review of fuel pre-injection in
39 105
flow, Trans. Jpn. Soc. Aeronaut. Space Sci. 46 (154) (2004) 217–223. supersonic, chemically reacting flows, Appl. Mech. Rev. 60 (4) (2007) 139–148.
40 [83] G.L. Wang, L.W. Chen, X.Y. Lu, Effects of the injector geometry on a sonic [117] D.P. Wang, Z.X. Xia, Y.X. Zhao, Q.H. Wang, B. Liu, Vortical structures of su- 106
41 jet into a supersonic crossflow, Sci. China Phys., Mech. Astron. 56 (2) (2013) personic flow over a delta-wing on a flat plate, Appl. Phys. Lett. 102 (2013) 107
42 366–377. 061911. 108
[84] A. Ben-Yakar, M.G. Mungal, R.K. Hanson, Time evolution and mixing character- [118] D.P. Wang, Y.X. Zhao, Z.X. Xia, Q.H. Wang, L.Y. Huang, Experimental investi-
43 109
istics of hydrogen and ethylene transverse jets in supersonic crossflows, Phys. gation of supersonic flow over a hemisphere, Chin. Sci. Bull. 57 (15) (2012)
44 Fluids 18 (2006) 026101. 1765–1771. 110
45 [85] L.L. Yuan, R.L. Street, J.H. Ferziger, Large-eddy simulations of a round jet in [119] D.P. Wang, Y.X. Zhao, Z.X. Xia, Q.H. Wang, Z.B. Luo, Flow visualization of su- 111
46 crossflow, J. Fluid Mech. 379 (1999) 71–104. personic flow over a finite cylinder, Chin. Phys. Lett. 29 (8) (2012) 084702. 112
[86] V. Viti, R. Neel, J.A. Schetz, Detailed flow physics of the supersonic jet inter- [120] C.H. Kim, I.S. Jeung, B. Choi, T. Kouchi, G. Masuya, Flowfield characteristics of
47 113
action flow field, Phys. Fluids 21 (2009) 046101. a hypermixer interacting with transverse injection in supersonic flow, AIAA J.
48 [87] C. Schaupp, R. Friedrich, Large-eddy simulation of a plane reacting jet trans- 50 (8) (2012) 1742–1753. 114
49 versely injected into supersonic turbulent channel flow, Int. J. Comput. Fluid [121] A.A. Shekarian, S. Tabejamaat, Y. Shoraka, Effects of incident shock wave 115
50 Dyn. 24 (10) (2010) 407–433. on mixing and flame holding of hydrogen in supersonic air flow, Int. 116
51 [88] S. Kawai, S.K. Lele, Large-eddy simulation of jet mixing in supersonic cross- J. Hydrog. Energy 39 (19) (2014) 10284–10292, http://dx.doi.org/10.1016/ 117
flows, AIAA J. 48 (9) (2010) 2063–2083. j.ijhydene.2014.04.154.
52 118
[89] D.M. Peterson, G.V. Candler, Hybrid Reynolds-averaged and large-eddy simu- [122] J.A. Schetz, L. Maddalena, S.K. Burger, Molecular weight and shock-wave ef-
53 lation of normal injection into a supersonic crossflow, J. Propuls. Power 26 (3) fects on transverse injection in supersonic flow, J. Propuls. Power 26 (5) 119
54 (2010) 533–544. (2010) 1102–1113. 120
55 [90] F.W. Spaid, E.E. Zukoski, A study of the interaction of gaseous jets from trans- [123] A.N. Hakim, S. Aso, Y. Tani, Experimental study on effects of shock wave im- 121
verse slots with supersonic external flows, AIAA J. 6 (2) (1968) 205–212. pingement on supersonic combustion, Mem. Fac. Eng., Kyushu Univ. 68 (1)
56 122
[91] S. Aso, S. Okuyama, M. Kawai, Y. Ando, Experimental study on the mixing (2008) 31–42.
57 phenomena in supersonic flows with slot injection, AIAA Paper 91-0016, 1991. [124] S.H. Won, I.S. Jeung, B. Parent, J.Y. Choi, Numerical investigation of trans- 123
58 [92] W. Huang, W.D. Liu, S.B. Li, Z.X. Xia, J. Liu, Z.G. Wang, Influences of the turbu- verse hydrogen jet into supersonic crossflow using detached-eddy simulation, 124
59 lence model and the slot width on the transverse slot injection flow field in AIAA J. 48 (6) (2010) 1047–1058. 125
supersonic flows, Acta Astronaut. 73 (2012) 1–9. [125] A.M. Tahsini, S.T. Mousavi, Investigating the supersonic combustion efficiency
60 126
[93] A.T. Sriram, J. Mathew, Numerical simulation of transverse injection of circular for the jet-in-crossflow, Int. J. Hydrog. Energy 40 (2015) 3091–3097.
61 jets into turbulent supersonic streams, J. Propuls. Power 24 (1) (2008) 45–54. [126] E. Erdem, S. Saravanan, J. Lin, K. Kontis, Experimental investigation of trans- 127
62 [94] D.P. Rizzetta, Numerical simulation of slot injection into a turbulent super- verse injection flowfield at Mach 5 and the influence of impinging shock 128
63 sonic stream, AIAA J. 30 (10) (1992) 2434–2439. wave, in: 18th AIAA/3AF International Space Planes and Hypersonic Systems 129
64 [95] G. He, Y. Guo, A.T. Hsu, The effect of Schmidt number on turbulent scalar and Technologies Conference, Tours, France, 2012, AIAA Paper 2012-5800. 130
mixing in a jet-in-crossflow, Int. J. Heat Mass Transf. 42 (1999) 3727–3738. [127] W. Huang, Numerical prediction on the interaction between the incident
65 131
[96] E.M. Ivanova, B.E. Noll, M. Aigner, A numerical study on the turbulent Schmidt shock wave and the transverse slot injection in supersonic flows, Aerosp. Sci.
66 numbers in a jet in crossflow, J. Eng. Gas Turbines Power 135 (2013) 011505. Technol. 28 (2013) 91–99. 132
JID:AESCTE AID:3543 /REV [m5G; v1.171; Prn:11/01/2016; 13:33] P.13 (1-13)
W. Huang / Aerospace Science and Technology ••• (••••) •••–••• 13

1 [128] W. Huang, J.G. Tan, J. Liu, L. Yan, Mixing augmentation induced by the interac- [131] B. Choi, K. Takae, T. Kouchi, G. Masuya, Turbulent characteristics for jet in- 67
2 tion between the oblique shock wave and a sonic hydrogen jet in supersonic jected into supersonic flow with pseudo shock wave, J. Propuls. Power 28 (5) 68
3 flows, Acta Astronaut. 117 (2015) 142–152. (2012) 971–981. 69
[129] M.B. Gerdroodbary, D.D. Ganji, Y. Amini, Numerical study of shock wave in- [132] C.P. Goyne, J.C. McDaniel, T.M. Quagliaroli, R.H. Krauss, S.W. Day, Dual-mode
4 70
teraction on transverse jets through multiport injector arrays in supersonic combustion of hydrogen in a Mach 5, continuous-flow facility, J. Propuls.
5 crossflow, Acta Astronaut. 115 (2015) 422–433. Power 17 (6) (2001) 1313–1318. 71
6 [130] W. Huang, Z.G. Wang, M. Pourkashanian, L. Ma, D.B. Ingham, S.B. Luo, J. Lei, [133] S.L.N. Desikan, R. Saravanan, S. Subramanian, A.E. Sivararamakrishnan, S. Pan- 72
7 J. Liu, Numerical investigation on the shock wave transition in a three- dian, Investigation of supersonic jet interaction with hypersonic cross flow, 73
8 dimensional scramjet isolator, Acta Astronaut. 68 (2011) 1669–1675. J. Fluids Eng. 137 (2015) 101101. 74
9 75
10 76
11 77
12 78
13 79
14 80
15 81
16 82
17 83
18 84
19 85
20 86
21 87
22 88
23 89
24 90
25 91
26 92
27 93
28 94
29 95
30 96
31 97
32 98
33 99
34 100
35 101
36 102
37 103
38 104
39 105
40 106
41 107
42 108
43 109
44 110
45 111
46 112
47 113
48 114
49 115
50 116
51 117
52 118
53 119
54 120
55 121
56 122
57 123
58 124
59 125
60 126
61 127
62 128
63 129
64 130
65 131
66 132

You might also like