You are on page 1of 48

J. Fluid Mech. (2015), vol. 780, pp. 226–273.

c Cambridge University Press 2015 226


doi:10.1017/jfm.2015.454

Ignition, flame structure and near-wall burning


in transverse hydrogen jets
in supersonic crossflow

Mirko Gamba1, † and M. Godfrey Mungal2,3


1 Department of Aerospace Engineering, University of Michigan, Ann Arbor, MI 48109, USA
2 Mechanical Engineering Department, Stanford University, Stanford, CA 94305, USA
3 School of Engineering, Santa Clara University, Santa Clara, CA 95053, USA

(Received 10 October 2014; revised 30 May 2015; accepted 30 July 2015;


first published online 3 September 2015)

We have investigated the properties of transverse sonic hydrogen jets in high-


temperature supersonic crossflow at jet-to-crossflow momentum flux ratios J between
0.3 and 5.0. The crossflow was held fixed at a Mach number of 2.4, 1400 K and
40 kPa. Schlieren and OH∗ chemiluminescence imaging were used to investigate
the global flame structure, penetration and ignition points; OH planar laser-induced
fluorescence imaging over several planes was used to investigate the instantaneous
reaction zone. It is found that J indirectly controls many of the combustion processes.
Two regimes for low (<1) and high (>3) J are identified. At low J, the flame is
lifted and stabilizes in the wake close to the wall possibly by autoignition after some
partial premixing occurs; most of the heat release occurs at the wall in regions where
OH occurs over broad regions. At high J, the flame is anchored at the upstream
recirculation region and remains attached to the wall within the boundary layer
where OH remains distributed over broad regions; a strong reacting shear layer exists
where the flame is organized in thin layers. Stabilization occurs in the upstream
recirculation region that forms as a consequence of the strong interaction between
the bow shock, the jet and the boundary layer. In general, this interaction – which
indirectly depends on J because it controls the jet penetration – dominates the
fluid dynamic processes and thus stabilization. As a result, the flow field may be
characterized by a flame structure characteristic of multiple interacting combustion
regimes, from (non-premixed) flamelets to (partially premixed) distributed reaction
zones, thus requiring a description based on a multi-regime combustion formulation.
Key words: turbulent reacting flows

1. Introduction
A jet in crossflow is one of the most fundamental canonical flows for the
investigation of turbulent mixing and combustion at both the fundamental and the
applied level. It is of relative simplicity, yet it maintains many features of interest,
such as three-dimensionality, separation and recirculation regions, wall-bounded

† Email address for correspondence: mirkog@umich.edu


Reacting transverse jet in supersonic crossflow 227
effects and vortical flow regions. Furthermore, it has a level of practical relevance
to engineering applications. A jet in crossflow can be found in many technological
processes, from the industrial level (e.g. an industrial flare stack) to fuel delivery
strategies in propulsion applications. A significant body of work has been conducted
to characterize and study this system, at both low and high (i.e. supersonic) speeds,
especially under non-reacting conditions. Two recent review articles on the study of
transverse jets are by Karagozian (2010) and Mahesh (2013).
Transverse jets in crossflow (using hydrogen or other hydrocarbon fuels) have
long been candidates as simple fuel delivery configurations in supersonic combustion
engines (Huber, Schexnayder & McClinton 1979). Arrays of single or staged injectors
have in fact been proposed and investigated as fuel delivery strategies for air-breathing
high-speed propulsion systems ranging from ramjet to scramjet engines. However, their
limitations in flame-holding, fuel penetration and mixing characteristics are known in
these applications; thus, other configurations that include various mixing-enhancement
strategies have been considered (Gutmark, Schadow & Yu 1995; Seiner, Dash &
Kenzakowski 2001) and continue to be developed. Some of the application-driven
configurations are based on wall injection in the presence of a cavity (Donbar
et al. 2000; Rasmussen et al. 2005; Rasmussen, Dhanuka & Driscoll 2007; Ryan
et al. 2009; Hsu et al. 2010) or hyper-mixing configurations where vorticity-induced
mixing is used to improve the overall performance and operability of supersonic
combustors (Hartfield, Hollo & McDaniel 1994; Gauba et al. 1997; Doster et al.
2009; Vergine et al. 2015).
Many of the general flow properties of a jet in crossflow are also found in the
supersonic counterpart – the jet in supersonic crossflow (JISCF). Low-speed transverse
jets in crossflow are characterized by a complex system of vortical structures generated
by the interaction of the transverse jet with the crossflow (Fric & Roshko 1994; Kelso
& Smits 1995; Smith & Mungal 1998). The near field is characterized by horseshoe
vortices that wrap around the base of the jet and by shear layer vortices that develop
in the windward side of the transverse jet. The far field is dominated by a persistent
counter-rotating vortex pair (CVP). The CVP forms in the near field by reorganization
of shear layer vorticity into a folded vortex sheet that deflects in the far field to
align (the axis of the CVP) with the direction of the crossflow (Kelso, Lim & Perry
1996; Cortelezzi & Karagozian 2001). Finally, for sufficiently strong transverse jets,
the wake of the jet is characterized by wake vortices that are generated near the wall
and stretch towards the deflected jet (Fric & Roshko 1994). The supersonic counterpart
retains some of these most important vortical structures. The main new contribution
to the flow field comes with the formation of a complex system of shocks due to the
supersonic crossflow (see the schematic diagrams of figure 1).
A typical configuration used in the supersonic regime is to generate the transverse
jet by underexpanded sonic injection at an angle (perpendicular, for example, as
we consider in this study) to the supersonic crossflow from a surface. A large
body of fundamental work has been directed at detailing the general structure and
performance of this system (Dowdy & Newton 1963; Schetz, Hawkins & Lehman
1967; Billig, Orth & Lasky 1971; Cohen, Coulter & Egan 1971; Povinelli & Povinelli
1971; Papamoschou & Hubbard 1993; Gruber et al. 1996), its flow field and mixing
characteristics (Rogers 1971b; Gruber et al. 1995; Santiago & Dutton 1997; Everett
et al. 1998; Gruber & Goss 1999; VanLerberghe et al. 2000; Kobayashi et al. 2007),
the effect of the boundary layer (Dowdy & Newton 1963; McClinton 1974) and
injection details (e.g. angle, staging, etc.) (Bier, Kappler & Wilhelmi 1971; McDaniel
& Graves 1988; Hollo, McDaniel & Hartfield 1994; Hartfield & Bayley 1995).
228 M. Gamba and M. G. Mungal

(a)
Bow shock Jet boundary

Upstream
recirculation
Mach disk

Separation
shock Downstream
recirculation

Jet fluid

(b) Shear layer

Bow
y shock

x
CVP
z Boundary
layer/wake
Horseshoe
vortex
Plane of
symmetry

Upstream
recirculation
Boundary
layer

Flat plate

F IGURE 1. (Colour online) Schematic diagrams of a transverse jet in supersonic crossflow


showing some of the dominant flow features: (a) side-view; (b) three-dimensional view
showing the transverse jet injected from a flat plate.

Similarly to underexpanded jets in quiescent environments (Crist, Sherman & Glass


1966; Schetz et al. 1967), transverse underexpanded jets form a complex system of
expansion and shock waves (referred to as the barrel shock) as the jet expands into
the crossflow. The barrel shock is deformed by the hydrodynamic forces resulting
from the interaction with the incoming crossflow. A shear layer forms at the boundary
between the deflected barrel shock and the crossflow. The flow blockage provided
by the jet induces a three-dimensional bow shock that wraps around the transverse
jet. For sufficiently large penetration of the jet, the bow shock is nearly normal at
the base and folds around the jet resulting in diminishing strength. The bow shock
stabilizes upstream of the jet and generates a rise in wall pressure upstream of the
jet. When this pressure rise is larger than three times the undisturbed (free stream)
value, the boundary layer separates (Dowdy & Newton 1963) forming the upstream
recirculation region which is characterized by two adjacent regions of swirling flow
Reacting transverse jet in supersonic crossflow 229
(Everett et al. 1998): an upstream bubble associated with the boundary layer followed
by a downstream one resulting from the presence of the jet. A three-dimensional
separation oblique shock is also generated as a response to the upstream separation
region. The separation and bow shocks intersect somewhere at the jet. The details of
the upstream separation region depend on the characteristics of the incoming boundary
layer (e.g. laminar or turbulent) and the size of the transverse jet (e.g. diameter). For
instance, the pressure rise in the separation region in a laminar boundary layer is lower
than in a turbulent boundary layer, although its length is larger by approximately a
factor of three (Dowdy & Newton 1963). Finally, because of the three-dimensionality
of the bow shock, the incoming boundary layer experiences a three-dimensional
interaction with the bow shock that extends further away from the immediate vicinity
of the injector itself.
Many parameters describing the incoming crossflow and the jet itself are found
to be important in describing the system. For example, the ratio of the boundary-
layer thickness to the jet exit diameter is an important quantity to describe near-field
events (McClinton 1974). The primary global flow-controlling parameter is the jet-to-
crossflow momentum flux ratio, defined as

ρj Uj2 pj γj Ma2j
J= = , (1.1)
ρa Ua2 pa γa Ma2a

where ρ and U are the fluid density and speed respectively, Ma is the Mach number,
p is the static pressure and γ is the ratio of specific heats. The subscripts ‘a’ and
‘j’ refer to crossflow and jet bulk conditions respectively. Normally, the conditions
describing the jet properties in (1.1) are taken at the jet exit. This parameter controls
the overall penetration and spreading of the JISCF. A modification to the above
relation based on the concept of an ‘effective back pressure’ can also be introduced
to describe the initial near-field structure of the barrel shock (such as its size, height
of the Mach disk and mass average properties along the centreline; Schetz & Billig
1966; Billig et al. 1971).
Because of the performance limitations associated with mixing in supersonic
combustion (Ferri 1973), mixing in the JISCF has been a primary focus of previous
work (Rogers 1971a,b; Torrence 1971; McDaniel & Graves 1988; Gruber et al.
1995; Santiago & Dutton 1997; VanLerberghe et al. 2000; Lin et al. 2010). Many
global morphological metrics such as the jet penetration have been defined and
studied in these mixing studies. These studies have identified the importance of
the windward shear layer on the entrainment process, along with the role of
the CVP in the entrainment process in the wake. Because of the important role
large-scale coherent structures play in the scalar transport processes, a large amount
of work has been dedicated to understanding their dynamics and interaction with the
supersonic crossflow (Gruber et al. 1996, 1997b; Ben-Yakar, Mungal & Hanson 2006).
Furthermore, the effect of compressibility (measured by an increase in convective
Mach number) on the evolution of large-scale structures and its implication for the
observed decrease in mixing have also been investigated (Gruber et al. 1997a). In
recent years, large-eddy simulation (LES) has been an important tool for better
understanding the mixing and dynamics of the JISCF (Génin & Menon 2010; Kawai
& Lele 2010; Peterson & Candler 2010).
Reacting transverse jets in the supersonic high-enthalpy regime have also received
attention. Previous work has primarily focused on developing diagnostic capabilities
for these flows (Allen et al. 1993; McMillin, Seitzman & Hanson 1994) and on
230 M. Gamba and M. G. Mungal
investigating the ignition and stability characteristics of hydrogen transverse jets
(Yoshida & Tsuji 1977; Lee et al. 1992; Rothstein & Wantuck 1992; Ben-Yakar &
Hanson 1998; Heltsley et al. 2007). More recently, work has also been directed at
investigating the reacting characteristics of more complex fuel injection approaches
that can be found in scramjet applications (Donbar et al. 2000; Rasmussen et al. 2007;
Ryan et al. 2009). More recently, reacting transverse jets in subsonic crossflow have
also found renewed interest (Huang & Chang 1994; Grout et al. 2011; Kolla et al.
2012; Micka & Driscoll 2012; Schaupp, Friedrich & Foysi 2012; Sullivan et al. 2014).
In the current study, we consider a single isolated transverse hydrogen jet injected
perpendicularly to a high-enthalpy supersonic crossflow characteristic of Mach 8
flight enthalpy (pressure of 40 kPa, temperature of 1400 K and Mach number of
2.4) generated with an expansion tube. The objective of the work is to investigate
the effect of the momentum flux ratio on ignition, stability, and global and local
combustion characteristics of the system at fixed free stream conditions. We make
use of OH planar laser-induced fluorescence (PLIF) to mark the instantaneous
reaction zone over several planes along the principal orthogonal directions (primarily
streamwise/wall-normal and streamwise/spanwise planes). This allows us to generate
a detailed view of the three-dimensional reacting flow field. We particularly focus on
analysing the instantaneous structure of the reacting flow near the wall, where most
of the reaction and flame stabilization is observed to take place.
In the following sections, a description of the experimental configuration and flow
diagnostics will first be given, followed by a description of the main observations on
the flow field and flame structure. The importance of the flow features resulting from
the interaction of the transverse jet with the supersonic crossflow on ignition, flame
stabilization and flame structure will be discussed in the final sections.

2. Experimental facility and configuration


In this section, a description of the experimental facility, flow set-up and imaging
configuration is given. A detailed description of the characterization of the flow
condition produced by the expansion tube is reported in appendix A.

2.1. The expansion tube


The experiments presented here were conducted at the 600 Expansion Tube Facility
of the High Temperature Gas Dynamics Laboratory at Stanford University, which
was used to generate the high-enthalpy supersonic crossflow used in the experiments.
A schematic diagram of the expansion tube is shown in figure 2. An account of
the design and operation of this facility is given by Heltsley et al. (2006); recent
modelling efforts for this facility have been provided by Örley et al. (2011, 2012). The
operational principle and a theoretical analysis of the performance of an expansion
tube to generate high-enthalpy supersonic and hypersonic flows can be found in
the seminal work of Trimpi (1962). In brief, the expansion tube takes advantage
of the unsteady shock and expansion waves produced in the tube to generate a
short-duration (∼1 ms) supersonic high-enthalpy flow in the test section. The facility
provides optical access for flow field imaging experiments.
Calibration and characterization of the flow conditions generated by the expansion
tube was performed using different approaches. All the results of these activities
are described in appendix A. In brief, the calibration measurements established
the following bulk-average free stream conditions for the experiment: Maa = 2.4,
pa = 40 kPa and Ta ≈ 1400 K (with a corresponding stagnation temperature of
approximately 3000 K). The useful test time was approximately 500 µs.
Reacting transverse jet in supersonic crossflow 231

Driver
section
Primary
From laser Secondary Driven diaphragm
source diaphragm section

Expansion
Test section
section
Sheet-forming To driven section
optics shock counters

Dump To expansion section


tank shock counters
ICCD
camera

F IGURE 2. (Colour online) Schematic diagram of the expansion tube showing the main
sections, diaphragm assemblies, shock speed counters, and the flow laser illumination and
imaging arrangement. The test section offers optical access on three sides.

2.2. Transverse jet


The transverse jet was generated by issuing fuel perpendicularly to the crossflow from
a flat plate. A three-dimensional schematic diagram of the flat plate configuration,
along with some of the relevant flow features is shown in figure 1(b). The flat plate
was composed of an aluminium plate with a hardened A-2 tool steel sharp leading
edge (hardening of the leading edge was required to prevent damage during operation
of the expansion tube). The resulting flat plate was approximately 155 mm (6.125 in.)
long and 100 mm (4 in.) wide. The flat plate was mounted in the test section within
the region of uniform and undisturbed supersonic flow bounded by the limiting
characteristic region of the incoming flow and its inviscid core. A more detailed
description of the flow characteristics of the incoming supersonic flow, including an
assessment of temporal and spatial uniformity, is presented in appendix A.
Ultra-high-purity grade (99.999 % purity) hydrogen was used as fuel throughout
the study. Fuel was injected from a contoured round injector with a d = 2 mm
diameter and located on the centreline of the plate at approximately 64 mm from
the leading edge. Fuel was delivered from a high-pressure gas cylinder, regulated by
a pressure regulator, and delivered to the injector through high-pressure tubing. Fuel
injection was controlled by a fast-acting solenoid valve (Valvetech model 15060-18)
directly mounted beneath the flat plate at the plenum of the injector. To ensure a
fully established flow to the injector, fuel injection was initiated at the beginning of
the experiment well before the onset of the flow in the test section (see appendices A
and B for further details). No active ignition device was used to ignite and stabilize
the flame.
Six different jets generated at different jet stagnation pressures, hpj,o i, under the
same crossflow conditions were considered in the study. Some details about the
calibration of the injector and the injection timing relative to the short-duration test
are given in appendix B. The set of cases is summarized in table 1 and spans a range
of momentum flux ratio, J, from 0.3 to 5.0 at fixed free stream conditions. The values
of J were computed according to the definition (1.1) assuming isentropic expansion
from hpj,o i to sonic conditions and pa = 40 kPa. It should be noted that the value of
hpj,o i reported in table 1 refers to the test time-averaged value of pj,o extracted from
232 M. Gamba and M. G. Mungal

hpj,o i Uncertaintya J Red


(kPa) (%) (—) (—)
119 12.3 0.29 18 800
289 4.0 0.71 45 800
459 2.1 1.14 72 800
723 1.5 1.79 114 600
1098 1.7 2.71 174 000
2034 1.2 5.03 322 400
TABLE 1. Summary of the hydrogen jet injection parameters: J was computed assuming
pa = 40 kPa; Red was computed based on the cold flow properties of hydrogen at the jet
exit plane assuming choked conditions.
a
Uncertainty refers to hpj,o i.

the calibration step. The jet exit Reynolds number, Red = ρj uj d/µj , based on the cold
properties at the jet exit plane is also reported. The indicated uncertainty in table 1
refers to the uncertainty in hpj,o i and accounts for both variation during the test time
and measurement (calibration) uncertainty.
A direct quantification of the characteristics of the incoming boundary layer that
develops on the flat plate was not possible. Schlieren imaging of the incoming
flow (see later) suggests that the boundary layer remains thin, on the scale of
approximately half jet diameter. Therefore, analytical results for supersonic laminar
boundary layers (White 1991) were used to estimate the boundary-layer thickness of
the incoming flow. Under the free stream conditions of this study and assuming an
isothermal wall near room temperature, the boundary-layer thickness was estimated
to be approximately 1 mm. The unit Reynolds number of the incoming flow was
approximately 3 × 106 m−1 and the local Reynolds number at the injector location
was approximately 0.2 × 106 . Under these conditions, a laminar boundary layer would
be expected (White 1991).

2.3. Imaging configuration


In the study, different imaging techniques were used to investigate the system: the
flow and shock structure was visualized by Schlieren, the overall characteristics of
the reacting jet were investigated by OH∗ chemiluminescence imaging, whereas the
instantaneous reaction zone was visualized by OH PLIF imaging.
The Schlieren system used in the study was a regular folded Z-type set-up. The
illumination source was provided by a long-duration xenon flashlight, two 2 m focal
length spherical mirrors were used to collimate and refocus the illumination source on
the knife edge, and a LaVision ImagePro camera (resolution 1376 × 1040 pixel) was
used for imaging. Background subtraction and normalization were performed on the
acquired images. Instantaneous (500 ns exposure) and time-averaged (20 µs exposure)
single-event images of the jet structure were acquired.
For the OH PLIF imaging experiments, the Q1 (N 00 = 7) line near 283.3 nm of
the A2 Σ + ← X 2 Π(v 0 = 1, v 00 = 0) electronic transition was selected. Broadband
fluorescence from the whole upper excited state to all lower states was collected. The
excitation source was provided by the frequency-doubled output of a tunable dye
laser (Lumonics HD-500) pumped by a frequency-doubled Nd:YAG laser operated
at 10 Hz (Continuum Powerlight 8000). The energy delivered was approximately
Reacting transverse jet in supersonic crossflow 233
10 mJ per pulse. As shown schematically in figure 2, the excitation laser source
was reflected towards the test section of the expansion tube and formed into a thin
sheet by a combination of cylindrical lenses. The laser sheet thickness was estimated
to be of the order of 300 µm. A gated intensified CCD (ICCD) camera (Andor
iStar, resolution 512 × 512 pixels) fitted with a Nikon 105 mm, f /4 UV lens was
placed perpendicularly to the illumination plane to collect fluorescence. UG-11 and
WG-295 filters were used to reject unwanted light. The ICCD camera was gated to
400 ns to prevent interference of background illumination. Background subtraction
and (mean) laser sheet non-uniform intensity correction was applied to the imaged
fields. Laser energy shot-to-shot variation was also corrected for by monitoring the
energy delivered during each imaging experiment. However, no attempt was made to
correct for shot-to-shot variation of the laser sheet non-uniformity. Several side- and
plan-view planes were considered. With reference to the coordinate system defined in
figure 1(b), side-view planes are defined to be streamwise/wall-normal (x–y) planes,
whereas plan-view planes are defined to be streamwise/spanwise (x–z) planes.
The OH∗ chemiluminescence imaging experiments used an imaging configuration
very similar to the OH PLIF experiments and differed only in the choice of spectral
filter. For this case, a bandpass interference filter centred at 313 nm with a full-width
at half-maximum of 10 nm was used on a similar ICCD camera configuration to
that used for the OH PLIF imaging. Only long-exposure (200 µs) imaging was
considered to obtain a time-averaged and line-of-sight-integrated representation of the
overall location of the burning regions. In this work, the OH∗ chemiluminescence
is used as an approximate indicator of the global heat release distribution in the
flow field. No calibration to relate heat release rate to emission intensity has been
attempted and the reported results are used to draw conclusions based on a relative
scale. Therefore, in order to allow for a direct comparison between different cases,
the imaging configuration (such as magnification, field of view, etc.) and camera
settings (such as gain on the intensifier, camera lens aperture, etc.) were maintained
unaltered between all cases considered in the study. A discussion on the use of OH∗
chemiluminescence emission as a heat release rate indicator will be presented below.
To correctly interpret some of the results discussed below on the instantaneous
structure of the reaction zone, it was necessary to describe the resolution of the
imaging system used in the OH PLIF experiments. Therefore, the modulation transfer
function (MTF) of the imaging system was approximately constructed using the
scanning knife-edge technique (Clemens 2002). A uniform white field was generated
by an LED array placed behind a stack of ground glass plates and it was imaged
with the ICCD camera as the knife edge was scanned across the field of view, from
which the step-response function (SRF) was constructed at discrete points in the field
of view. It was found that the SRF was well approximated by an error function, in
which case the MTF is Gaussian. Knowledge of the MTF then allowed us to correct
the length scales extracted from the OH PLIF imaging following the method outlined
by Wang & Clemens (2004).
Due to the short duration of the available test time, all of the imaging experiments
reported in this work are single-shot realizations repeated under nominally identical
run conditions. Some of the composite figures showing the instantaneous flame
structure reported in the following sections have been assembled from uncorrelated
realizations. Furthermore, all of the single-shot imaging was carried out at the same
nominal time delay with respect of the initiation of the flow (i.e. to the master
trigger), well after a stationary flow was obtained. By taking OH PLIF images
at different time delays with respect to the master trigger, a (pseudo) time series
234 M. Gamba and M. G. Mungal
of the instantaneous reaction zone was constructed (not shown) to assess transient
effects leading to the stationary flow and to ensure that the remaining data were
taken under stationary flow conditions. Similarly, kilohertz-rate OH∗ movies were
taken to further assess the ignition transient. With reference to the pressure trace
of figure 21 (shown in appendix A), the steady portion (of the pressure time trace)
during the test time begins approximately 4.75 µs from the master trigger, but a
fully established stationary jet and flame structure is not observed until 4.9 ms. All
imaging reported here has therefore been acquired well past this transient time. Thus,
the OH PLIF and Schlieren imaging was fixed at 5.1 ms, whereas the long-exposure
OH∗ chemiluminescence imaging was initiated at 5.0 ms with a 200 µs exposure.

3. Global structure and features of the reacting transverse jet


In this section the global characteristics of the reacting transverse jets are presented.
First, the overall jet and shock structure as visualized by Schlieren imaging will be
briefly presented. Then, the global burning and ignition characteristics as well as the
far-field flame penetration and heat release distribution will be investigated from OH∗
chemiluminescence imaging. The detailed instantaneous flame structure as marked by
the OH radical will be investigated in the following section. The general flow and
shock structure of a JISCF, and its dependence on the working parameters (such as
J), is now well known, as discussed earlier. However, because of its importance for
the flame and stabilization characteristics that will be discussed below, a summarizing
description for the particular cases under investigation is given first.

3.1. Shock structure


The structure of the shock system for the case of the current experiments is shown
in figure 3 where instantaneous (500 ns exposure) and time-average (20 µs exposure)
Schlieren images are presented. The main characteristics have been outlined in the
schematic diagram of figure 1 and correspond to the labelled features indicated
in figure 3(a). In all figures presented here, the crossflow is from left to right.
Figure 3(a) shows an instantaneous snapshot for the J = 5.0 case. The image captures
the incoming flow as it develops over the flat plate (indicated by the black outline)
and the near field of the transverse jet. Well upstream of the injector, two shock wave
systems are observed. First, a lip shock is formed by the interaction of the incoming
supersonic flow with the boundary layer developing on the flat plate; second, a set
of oblique waves is observed to originate from the flat plate (near x/d = −15). The
second shock wave system does not result from injection, nor does it originate by
the mating of the sharp leading edge and the plate (which is near x/d = −5). Its
nature and origin could not be identified. However, it does not appear to interact
with the transverse jet; thus, it is not expected to affect the jet. Figure 3(b) shows
the corresponding time-average Schlieren image for the same case as figure 3(a). A
few jet diameters upstream of the injector, the near field and part of the wake of
the transverse jet are shown. Finally, figure 3(c,d) shows the time-averaged shock
structure generated around the J = 1.8 and J = 0.3 cases.
The salient JISCF features that the Schlieren images identify are the bow shock
(B) curving around the jet, the separation oblique shock (C), the deformed barrel
shock (A) terminating in a tilted Mach disk, the recompression shock (E) at the
lee side of the transverse jet following reattachment and the shear layer of the jet
(F). Furthermore, these characteristics are strongly influenced by the value of J. In
particular, the penetration of the barrel shock increases with J; similarly, the deflection
Reacting transverse jet in supersonic crossflow 235

(a) 10
Barrel shock Bow shock Shear layer
Lip shock
Recompression
Separation shock shock
Unidentified
5 disturbances Recirculation

0
–30 –25 –20 –15 –10 –5 0 5 10
(b) 10
B
F
5 C E
A
D
0
–10 –5 0 5 10 15 20 25 30

(c) 10
B

5
C F
G
A
0
–10 –5 0 5 10 15 20 25 30
(d) 10
B

5
G
F
0

–10 –5 0 5 10 15 20 25 30

F IGURE 3. Schlieren images of the flow and shock structure around transverse
underexpanded jets in supersonic crossflow: (a) instantaneous (500 ns exposure) Schlieren
image for J = 5.0, (b) time-averaged (20 µs exposure) Schlieren image for J = 5.0,
(c) J = 1.8 and (d) J = 0.3; A, barrel shock; B, bow shock; C, separation shock; D,
upstream separation region; E, recompression shock; F, shear layer; G, Mach wave.

and deformation of the barrel shock system is accentuated as J is lowered until it


disappears as it merges with the wake and boundary layer, as for the J = 0.3 case,
or until the subsonic limit is reached. The penetration of the jet into the crossflow
provides flow blockage to the incoming flow, hence generating a three-dimensional
bow shock whose strength depends on the level of blockage (i.e. the value of J).
The role and importance of the bow shock on the combustion characteristics will
be emphasized further below when the global and instantaneous flame structure is
discussed. One aspect to note here is the importance of the interaction of the bow
236 M. Gamba and M. G. Mungal
shock with the incoming flow, especially with the boundary layer. As a result of
the bow-shock/boundary-layer interaction, the incoming boundary layer separates (a
process that is favoured with a laminar boundary layer; Dowdy & Newton (1963),
Rogers (1971a,b)) and a recirculation region forms in front of the jet (D). The
recirculation region is visible in the J = 5.0 case, but it is a fairly small feature in
the J = 1.8 and 0.3 cases. The strength and size of this flow feature are related to the
strength of the interaction, which is controlled by the blockage-induced bow shock,
hence J. As a result of the recirculation region, a separation shock forms in front of
the separation bubble (see figure 3a). Both features play an important role in flame
stabilization and flame holding (Dowdy & Newton 1963; Schetz & Billig 1966; Schetz
et al. 1967; Huber et al. 1979; Papamoschou & Hubbard 1993; Gruber et al. 1996;
Ben-Yakar & Hanson 1998). The outline of the wake of the jet is visible downstream
of the injection point and, past the recompression shock, it is not significantly affected
by further shocks. In fact, the effect of the shock system generated by transverse
injection is limited only in the near-field region as the flow is deflected by the
jet, whereas the wake of the jet is relatively shock-free. This also implies that the
pressure distribution in the wake of the (isolated) transverse jet quickly recovers to
the undisturbed value, as surface pressure measurements have demonstrated (Everett
et al. 1998; Gruber & Goss 1999). The details pointed out in this section have a
role in the interpretation of the results drawn from the reaction zone imaging given
further below; thus, they will be discussed further in the following sections. It should
be noted that the weak wave labelled G in figure 3(c,d) is not associated with the
separation shock, but it is a Mach wave emanating from the line joining the leading
edge and the flat plate located at x/d = −5. This Mach wave is visible only for cases
with J < 3 since the length of the upstream recirculation region is less than 5d; thus
the flow can respond to this geometric feature by creating a Mach wave.

3.2. Global characteristics of the burning transverse jet


The overall ignition and flame characteristics of the system were investigated
using OH∗ chemiluminescence imaging from naturally occurring chemically excited
hydroxyl radicals as they decay from the A2 Σ + excited electronic state to the X 2 Π
ground state through spontaneous radiative emission. The emitting excited species is
generated by the chemical reaction process and it is limited by collisional quenching.
Its emission intensity is related to the amount of the excited species. In premixed
flames, for constant equivalence ratio, unburnt mixture temperature and heat losses,
chemiluminescence has been shown to be an approximate direct indicator of the
global heat release rate. However, its interpretation as an indicator of heat release
rate is complicated by its dependence on turbulent fluctuations, strain rates, flame
surface curvature and a weak dependence on unmixedness (Lee & Santavicca 2003;
Ayoola et al. 2006). Because of these dependences, relating the emission intensity to
the heat release rate is only approximate, especially in the context of non-premixed
combustion as considered in the current study. Thus, chemiluminescence imaging is
used in this study only for a qualitative description of the burning characteristics,
their morphology and dependences. It should be noted that the approach followed
in the present study provides a line-of-sight and time-integrated representation of
the combusting flow field from a lateral view. Furthermore, multiple repetitions were
compared to ensure repeatability in the average characteristics.
Examples of the time-averaged OH∗ chemiluminescence images (time integration
of 200 µs) are shown in figure 4 for transverse jets at (a) J = 0.3, (b) J = 1.8 and
Reacting transverse jet in supersonic crossflow 237

(a) 15

0 0.5 1.0
10

A
5

0
–10 –5 0 5 10 15 20 25 30 35 40 45

(b) 15

0 0.5 1.0
10

0
–10 –5 0 5 10 15 20 25 30 35 40 45

(c) 15

0 0.5 1.0
10
B

5
C
D
0
–10 –5 0 5 10 15 20 25 30 35 40 45

F IGURE 4. Line-of-sight integrated time-averaged OH∗ chemiluminescence imaging of the


overall flame structure: (a) J = 0.3, (b) J = 1.8 and (c) J = 5.0. For all cases the exposure
was set to 200 µs. The labels and lines are described in the text.

(c) J = 5.0. The colourmap shown is linear and ranges from blue in the non-reacting
regions to red and yellow in the reacting ones and terminates with white at the
maximum intensity. To ensure that a meaningful comparison between the different
cases can be made, the OH∗ imaging configuration and data reduction steps were the
same for all injection cases, as also explained in § 2.3. Furthermore, the colourmap
scale used in constructing figure 4 is the same for all cases presented here.
The integration time used in the OH∗ imaging experiments corresponds to more
than 20 times the characteristic large-scale convective time scale, τc , of the system,
which ensures convergence to a stationary (and shot-to-shot repeatable) representation
of the combusting regions. The convective time scale τc is defined as yo /U, which
is the ratio between a characteristic large-scale length scale and the velocity scale of
the system. The characteristic length scale yo is taken as the maximum jet penetration
(for example, from the results of figure 3) observed within the field of view; the
characteristic velocity scale U is taken as the arithmetic average between the free
stream velocity (Ua = 1775 m s−1 ) and the jet exit velocity (Uj = 1250 m s−1 ). The
choice of this velocity scale is motivated by the observation that the convective
velocity of large-scale structures in JISCFs varies from Uj near the injector to
238 M. Gamba and M. G. Mungal
approximately Ua as the wake-like region of the JISCF is approached (Gruber et al.
1997a; Ben-Yakar et al. 2006).
Inspection of the set of images reveals that the characteristics of the overall burning
regions, their structure and the stabilization points depend on the value of J. The cases
reported in figure 4 are selected to demonstrate the limiting behaviours that the study
identifies. In particular, for values of J < 1 (figure 4a), it is observed that ignition
does not occur at the injection point, but the flame is stabilized at some downstream
point, which we refer to as the stabilization point xs , and it is indicated by the label
A in figure 4(a) (xs will be defined formally further below). Furthermore, the burning
region appears to be confined at the wall in a layer approximately 2d thick. As the
Schlieren image of figure 3(c) qualitatively indicates, the J = 3 case results in minimal
fuel penetration into the crossflow (of the order of the boundary-layer thickness), and
the wake of the jet remains attached to the wall.
As the value of J is increased, the jet penetration increases and the stabilization
point moves upstream. Nevertheless, the most intense burning region, as indicated
by intense OH∗ emission, still remains confined to the near-wall wake region of the
transverse jet. As the value of J is raised above unity (see figure 4b for the J = 1.8
case), weak shear layer burning is detected by the OH∗ imaging and exists above
the intense near-wall wake burning. For the J = 1.8 case, near-field burning on the
windward side of the transverse jet just behind the bow shock is also detected, but it
is a weak feature.
Burning of the shear layer, either in the near field in the region of maximum
deflection or downstream in the wake of the jet, becomes important only as J is
increased to values larger than ∼3. As the value of J is increased even further, the
downstream stabilization point and the near-field burning merge and the reacting
transverse jet is stabilized directly at the injector and upstream of it (see figure 4c).
However, even for the large value of J considered in figure 4(c), the near-wall
wake remains the region with the most intense burning as identified by the OH∗
emission. Furthermore, also for the J = 5.0 case, there exists a region of low
OH∗ emission downstream of the region of maximum deflection separating the
near-field, post-bow-shock burning and the near-wall-stabilized wake combustion.
This low-OH∗ -intensity region is interpreted as being the result of quenching of
the shear layer due to the rapid expansion of the transverse jet which decreases
its temperature and pressure, while it accelerates the flow and possibly increases
straining of the mixing region. As a result, the rates of chemical reaction might
decrease overall in this region until further downstream (x/d > 20) where temperature
and pressure recover, and combustion can progress favourably. For J larger than ∼3,
burning upstream of the injection point is observed (see figure 4c). This feature has
typically been associated with the upstream recirculation region which ignites and
serves as a point of stabilization of the combustion process in the JISCF system
(Huber et al. 1979) and has also been observed in previous work on reacting JISCF
(Lee et al. 1992; Ben-Yakar & Hanson 1998).
As a concluding remark, it is important to point out that in all cases considered
in the study most of the OH∗ chemiluminescence emission is detected close to the
wall in the wake of the transverse jet, and, on average, emission in the shear layer
is weaker, at least within the imaging region (up to 45 jet diameters downstream).
Therefore, under the assumption that OH∗ emission tracks the heat release distribution,
it appears that the region where most of the initial heat release occurs might be
confined in the wake of the jet near the wall.
Reacting transverse jet in supersonic crossflow 239
3.3. Flame lift-off

Assuming that the relative OH luminosity can be taken as a marker of heat release,
the results of figure 4 suggest that heat release might be negligible or completely
absent in the near-field region for low values of J (say J < 1). Thus, low-J reacting
transverse jets appear to be lifted similarly to low-speed jet and transverse jet flames.
We define the stabilization point xs as the point downstream from the injector location
where the first occurrence of OH∗ luminosity is observed. In extracting this quantity,
to be consistent among all cases, the OH∗ images of figure 4 were first low-pass
filtered (with a moving average filter of size 3 × 3 pixels) and then thresholded to
10 % of the maximum value of the OH∗ signal observed in the full data set. The
stabilization point xs is equivalent to the lift-off height or distance in low-speed flames.
Depending upon the mechanisms responsible for stabilization at xs (discussed below),
this distance could also be interpreted as an ignition delay distance (or time).
The stabilization point xs /d as a function of J is shown in figure 5. Three different
regions are identified in the figure and reflect the characteristics identified so far: (I)
a region with a weak J dependence where stabilization occurs in the wake of the
jet, near the wall (J < 1), (II) a region where mixed wake and near-field shear layer
ignition and stabilization occurs, and (III) a region where ignition and stabilization
occurs upstream of the jet and a flame is maintained at all downstream locations
(J & 3). Examples of the flame global structure consistent with each regime were
shown in figure 4. In regime I, xs is a weak function of J and approaches a limiting
value of approximately 22d as J → 0. Values of J < 0.3 were not investigated,
therefore it is not known whether blow-off occurs and at what value of J. For large
values of J in regime III, the reacting near-field shear layer, the upstream recirculation
region and the wake burning regions merge together. It seems reasonable to expect
that these features remain independent of J at larger values.
In figure 5 we also identify a second quantity of interest, which we refer to as the
location of heat release onset, xq , and, similarly to xs , it is reported as a function
of J. This quantity, which will be formally defined and discussed in more detail in
§ 3.5, indicates the point downstream of the injector where significant heat release
takes place and linearly decreases with J across the three regions (I, II and III)
identified in figure 5 from xs . In general, xq is different from (and larger than) the
stabilization point defined here; the latter defines where we first observe a stable
flame irrespective of how much heat is locally released at the stabilization point,
while the former defines where we first observe that a significant amount of heat has
been released as a result of the combustion process. We will return to this quantity
in § 3.5 when we discuss the distribution of heat release in the flow field.

3.4. Flame penetration


A second global characteristic is the penetration of the transverse jet flame into the
crossflow. In this work, we define the penetration of the reacting jet based on the OH∗
imaging results. Because OH∗ was not detected in the near field of the transverse
jet (for x/d < 20), penetration was defined only in the wake for x/d > 30 and it
was defined as the locus of points of the wall-normal location (y) at which the OH∗
intensity was 10 % of the maximum intensity value. Because all cases considered in
the study have the same distribution of values of OH∗ intensity, this definition is
consistent for all cases and with the definition of xs .
Figure 6(a) shows the extracted flame scaled penetration y/d as a function of the
scaled axial x/d coordinate after J-scaling. Penetration of jets in crossflow, in both
240 M. Gamba and M. G. Mungal

25

20
I
15

10
II
5

0
III
–5
0 1 2 3 4 5
J

F IGURE 5. The stabilization point (xs /d) and the location of heat release onset (xq /d) as
a function of the momentum flux ratio J.

(a)
J
0.3
0.7
1.1
10 1.8
2.7
5.0 (b) 1.2
Linear k
1.0
5 0.8
0.6
0.4 m
0.2
0 0
0 5 10 0 1 2 3 4 5
J

F IGURE 6. (a) Penetration of the flame boundary for the cases considered in the study
after J-scaling (3.1) with the parameters k and m left to depend on J. (b) Dependence of
m and k on J with best curve fit.

the subsonic (Margason 1993; Hasselbrink & Mungal 2001; Su & Mungal 2004) and
supersonic (Schetz & Billig 1966; Hersch, Povinelli & Povinelli 1970; Papamoschou
& Hubbard 1993; Gruber et al. 1995; Portz & Segal 2006) regimes, is found to be
primarily described by the parameter J, although weak effects on boundary-layer
thickness (McClinton 1974), crossflow Mach number (Portz & Segal 2006) and jet
fluid molecular weight (Ben-Yakar et al. 2006) might be present. In constructing
figure 6(a), it was assumed that the same holds for these reacting cases and for our
definition of flame penetration. Thus, the flame penetration was scaled according to
the following relation:  x m
y
=k . (3.1)
Jd Jd
In the above relation, the values of the premultiplicative factor k and of the exponent
m are allowed to vary with J and are computed as a least-squares fit of (3.1) to
the experimental observation. The J-scaled results are shown in figure 6(a). For the
Reacting transverse jet in supersonic crossflow 241
range of J values considered in the study, it is found that a relation of the type of
(3.1) describes the flame penetration well, but with a somewhat strong dependence
of k and m on J. It should be noted that in constructing the best fit of figure 6(a),
all possible parameters affecting penetration other than J (e.g. δ/d, Maa , molecular
weight) are effectively held constant in this study (δ is the boundary layer thickness
of the crossflow). Thus, if an extended model including these dependences were to
be used, extra dependences would primarily be reflected on k (Povinelli & Povinelli
1971; McClinton 1974; Portz & Segal 2006), but possibly also on m.
Figure 6(b) shows the dependence of the parameters k and m on J. It is found
that m varies from approximately 0.75 at low J and approaches a limiting value of
approximately 0.3 at large J; similarly, it is found that k varies from approximately
0.25 at low J to a limiting value of approximately 1 at large J. Furthermore, assuming
that these two parameters would follow a J scaling, best curve fits have been
constructed and are presented in this figure. In particular, the premultiplicative factor
is best correlated by k = k∞ (1 − eJqk ) with k∞ = 1.05 and qk ≈ −1, and the exponent
is best correlated by m = m∞ J qm with m∞ ≈ 1/2 and qm ≈ −1/4. These functional
forms are not supported on physical grounds, but are just proposed to conveniently
correlated these parameters within the limited range of J values considered in the
study. The variation of m and k is possibly a consequence of the range of low values
of J considered in the study, and they approach limiting values for J > 5.
The penetration of low-speed (subsonic) constant-density transverse jets have been
found to follow the form  x B
y
=A , (3.2)
rd rd

where r = J. This form has been successfully used to correlate a wide range of
experimental observations (e.g. Margason 1993) and it is consistent with conservation
of momentum scaling arguments (e.g. Hasselbrink & Mungal 2001). Typical values
for A and B are A ≈ 2 and B ≈ 0.3. To reflect the use of J, (3.2) can be written in
a form similar to (3.1) with k = AJ (B−1)/2 and B = m. Assuming that the low-speed
results apply to the supersonic regime, part of the variation of k with J could be
attributed to a direct dependence of J in the premultiplicative factor. However, the
data in figure 6(b) indicate ∂k/∂J > 0 whereas (3.2) implies ∂k/∂J < 0. Thus, the J
dependence of k found here has to be attributed to an intrinsic dependence possibly
due to the low value of J. It is also interesting to note that for our case, the exponent
tends to a limiting value of 0.3 for large J as in the subsonic case.
For the supersonic cases, the results of many studies support a functional form of
the type of (3.1). The results of some of these studies are summarized in table 2 where
the values of k, m and the conditions in those studies are reported. Previous studies
have covered a relatively broad range of non-reacting conditions, but they all reduce
to somewhat similar results. More notably, Povinelli & Povinelli (1971) studied the
effect of different parameters and found that a more comprehensive relationship was
the following (for x/d∗ > 7):
  x 0.204  θ 0.141
y po,jet 0.405 0.163
= 2.96 Maj + 0.5 , (3.3)
d∗ peb d∗ d∗
where d∗ is the throat diameter, θ is the boundary-layer momentum thickness and
peb is the effective back pressure (Billig et al. 1971). Their study, along with the
ones by McClinton (1974) and Portz & Segal (2006), emphasizes the importance of
the boundary-layer thickness on penetration and entrainment. In particular, McClinton
242 M. Gamba and M. G. Mungal

Reference k m Definition Range


Gruber et al. (1995) 1.23 0.344 90 % free stream 1<J<3
concentration Maa = 2
x/d < 5
He and CO2 into air
Rogers (1971b) 3.87J −0.557 0.143 0.5 % contour 0.5 < J < 1.5
Maa = 4.03
5 < x/d < 120
δ/d = 2.7
Rothstein (1992) 2.173J −0.276 0.281 OH PLIF layers J1
Maa = 1.5
x/d < 20
H2 into Ar/O2
Povinelli & Equation (3.3) — 0.5 % contour 0<J<1
Povinelli (1971) 2 < Maa < 3
1 < Maj < 4
0.2 < δ/d < 0.3
He into air
 0.0574
−0.557
δ
McClinton (1974) 4.20J 0.143 0.5 % contour J∼1
d Maa = 4.05
x/d = 120
1.25 < δ/d < 6.5
H2 into air
 0.221
−0.176
δ
Portz & Segal 1.362J 0.276 Schlieren 0.5 < J < 3
(2006)a d
Maa = 1.6 and 2.5
x/d < 30
0.8 < δ/d < 3.7
He, H2 , Ar into air
TABLE 2. Summary of some of the transverse jet penetration results from non-reacting
mixing studies. Unless otherwise noted, Maj = 1.
a
For simplicity, here the contribution of the molecular weight, Reynolds number and the
virtual origin term were neglected from the original formulation of Portz & Segal (2006).

(1974) found that δ/d has a weak effect on penetration defined as the 0.5 % jet
concentration contour (see table 2); on the contrary, a stronger effect was found
for jet penetration defined on the contour of local maximum concentration. In our
study, the crossflow is held fixed, thus the boundary-layer thickness dependence
cannot be assessed. Furthermore, our definition of penetration is more consistent
with a definition based on the 0.5 % contour (since we are tracking regions near the
stoichiometric contour), so we expect that overall the boundary layer would have a
secondary effect, at least for sufficiently large values of J (the work of McClinton
(1974) was carried out at a constant J = 1).
All of the previous work indicates a similar dependence of the premultiplicative
factor on J but no dependence on the exponent. Our result is somewhat different since
it gives the opposite dependence to what is typically observed. The discrepancy might
Reacting transverse jet in supersonic crossflow 243
be the result of the many different ways in which penetration has been defined and
the different range of conditions considered (Re, δ/d, M , Maa , x/d range, etc.).

3.5. Heat release distribution


The OH∗ chemiluminescence distributions shown in figure 4 are here used as a
proxy for the local heat release. Thus, the streamwise distribution of the heat release
was marked by computing the streamwise distribution of the time-averaged OH∗
chemiluminescence intensity, q(x/d), from the images of figure 4. Specifically, q(x/d)
was computed by adding together the measured values of the OH∗ signal at all pixels
along each column (i.e. for each one-pixel-wide column) in the image to generate a
one-dimensional streamwise profile (i.e. by integrating the 2D distribution along the
y-direction without any integration in the x-direction). This quantity was computed
in the same way for all cases considered here. The resulting profiles are shown in
figure 7(a). In general, we can observe a similar trend for all J cases considered.
For low-J cases where ignition does not occur at the injector itself, the distribution
of q(x/d) first increases far downstream from the injector and shows a continuous
increase up to a plateau; for the large-J cases, q(x/d) first peaks near the injector,
decreases immediately downstream of the injector (flame quenching) and then q(x/d)
increases back to a plateau. The point where the plateau is reached moves upstream
as J increases, and the value of the plateau increases with J.
The cumulative distribution of the heat release, Cq (x/d), was also computed and it
is shown in figure 7(b) for the different cases. After the initial energy release that
might exists directly at the injector, all J cases show that the majority of the heat
release occurs well downstream of the injection point with Cq (x/d) increasing linearly
with x/d after the initial region (this is a result of the plateau observed in q(x/d)).
From the profiles of Cq (x/d) we can also extract three more quantities: (i) xq , the
point of heat release onset; (ii) 1Cq , the slope of Cq (x/d) in the far field where the
profile becomes linear; and (iii) the total energy released within the imaging region
(i.e. the integral of Cq (x/d) over x/d). More specifically, the point of heat release onset
is defined as the intercept on the x-axis of the linear fit to Cq (x/d) in the far-field
linear portion – this definition is graphically shown in figure 7(b). One final quantity
extracted from the OH∗ visualizations is the peak value of the OH∗ signal, which in
all cases considered here occurs at a downstream location of approximately x/d = 30
near the wall at a wall-normal location of less than y/d = 1 (i.e. within what we can
consider the boundary layer).
Three of these quantities (peak OH∗ , 1Cq , total integrated q) are summarized
in figure 8 as a function of J, while xq is shown in figure 5 along with xs . It
should be noted that in constructing these quantities the OH∗ was normalized so
that the resulting quantities were of order 1 and could thus fit on the same scale. It
is interesting to note that the peak OH∗ (i.e. peak local heat release rate) and the
location where it occurs are the same for all cases considered as if the mechanism
responsible for it remains invariant upon J. As shown in figure 5, xq decreases
linearly with J from an extrapolated value of x/d = 25 as J → 0 to approximately
x/d = 17 for J = 5. Thus, although we observed that the point of first ignition (xs )
moves to the injector as J is increased, the point of heat release onset remains in
the far field. Thus, the use of strong injection (i.e. high J) can promote near-field
flame stabilization, but it has a much weaker impact on the point where the majority
of the heat release occurs (i.e. xq ). The result also implies that in the transverse jet
all of the heat release occurs far from the injector, thus requiring long combustion
chambers to achieve complete combustion.
244 M. Gamba and M. G. Mungal

(a) J (b)
1.0
0.3 1.5
0.7
1.1
1.8 1.0
q 0.5 2.7
5.0
0.5

0 0
0 10 20 30 40 0 10 20 30 40

F IGURE 7. (a) Normalized distribution of heat release q(x/d) (arbitrary units) and
(b) corresponding cumulative distribution Cq (x/d) for different J cases.

1.5

1.0

0.5

0
0 1 2 3 4 5
J

F IGURE 8. Parameters describing the heat release distributions of figures 4 and 7.

From the trends of the various quantities shown in figure 8, we can also observe
that there are two regions characterized by a different dependence of the heat release
distribution on J. For J 6 1, there is a strong nonlinear variation of the parameters
with J. On the contrary, for J > 1, the dependence is linear. Simple energy balance
considerations suggest that the cumulative heat releaser should increase linearly with
J (i.e. with the amount of fuel entering the system, thus J since Uj is constant).
This is observed only for J > 1. Therefore, for low J, other competing effects that
prevent rapid and uniform chemical reaction must exist. This is possibly related to the
stabilization process for low J since a similar dependence is found for xs (figure 5).

4. Instantaneous reaction zone structure


OH PLIF imaging is used to investigate the instantaneous structure of the reaction
regions on several side- and plan-view planes. Different off-centreplane side-view
planes and plan-view planes at different wall-normal heights from the flat plate were
considered, but only a selected set is shown here. The images shown subsequently
were constructed by appropriately combining results obtained from independent
single-shot imaging experiments over two adjacent fields of view, one capturing the
near field, the other showing the wake region. Thus, in some cases the reader will
observe a discontinuity (like a ‘break’) in the OH PLIF images; this discontinuity
is the seam between the individual images acquired at different times over adjacent
regions used to construct the combined images shown here.
Reacting transverse jet in supersonic crossflow 245

(a) 10

0
–5 0 5 10 15 20 25 30 35 40

(b) 10

0
–5 0 5 10 15 20 25 30 35 40

(c) 10

0
–5 0 5 10 15 20 25 30 35 40

F IGURE 9. Side-view OH PLIF images on the centreplane (z/d = 0) for (a) J = 0.3,
(b) J = 2.7 and (c) J = 5.0.

4.1. The reacting shear layer


Figure 9 shows an example of the instantaneous reaction zone marked by OH PLIF
for the symmetry centreplane (z/d = 0) of the transverse jet for three J cases: (a)
the low-value case of J = 0.3, (b) the intermediate-value case of J = 2.7 and (c) the
high-value case of J = 5.0. These three cases were selected to be representative of
the three regimes identified by the OH∗ imaging in figure 5. Figures 10 and 11 are
the corresponding OH PLIF images for off-centre planes located at z/d = 2 and 5
respectively. We observe that burning in the shear layer becomes stable (within the
available field of view) only when J is sufficiently large; in our case for J > 1.
Under the current flow conditions and for sufficiently high values of J (here J ∼ 5,
see figure 9c), three major flow features are observed: (i) an intermittently reacting
recirculation region upstream of the jet, (ii) a reactive shear layer on the windward
side of the transverse jet (figures 9c and 10c) and (iii) a highly reactive boundary
layer that extends laterally for several diameters off the jet centreplane (figure 11c).
Figure 11 clearly demonstrates the presence of the reactive boundary layer away from
the immediate vicinity of the wake of the jet and that this feature exists for all J
cases. At low values of J (figures 9a–11a), the shear layer does not sustain a stable
and continuous reaction layer, but the predominant burning region is in the wake near
the wall and away from the centreplane. For low J, ignition in the shear layer is
a relatively weak feature, even on the centreplane of the jet (figure 9a), and it is
not anchored in the near field. The burning upstream recirculation region is also not
observed. It seems therefore reasonable to assume that the intense near-wall burning
observed from OH∗ chemiluminescence is primarily due to the intense OH regions
246 M. Gamba and M. G. Mungal

(a) 10

0
–5 0 5 10 15 20 25 30 35 40

(b) 10

0
–5 0 5 10 15 20 25 30 35 40

(c) 10

0
–5 0 5 10 15 20 25 30 35 40

F IGURE 10. Side-view OH PLIF images on the plane z/d = 2 for (a) J = 0.3, (b) J = 2.7
and (c) J = 5.0.

(a) 10

0
–5 0 5 10 15 20 25 30 35 40

(b) 10

0
–5 0 5 10 15 20 25 30 35 40

(c) 10

0
–5 0 5 10 15 20 25 30 35 40

F IGURE 11. Side-view OH PLIF images on the plane z/d = 5 for (a) J = 0.3, (b) J = 2.7
and (c) J = 5.0.
Reacting transverse jet in supersonic crossflow 247

(a) 10

0
–5 0 5 10 15 20 25 30 35 40

(b) 10

0
–5 0 5 10 15 20 25 30 35 40

F IGURE 12. Average side-view OH PLIF image on the centreplane for (a) the J = 0.3
(nine-frame average) and (b) the J = 5.0 (11-frame average) JISCF.

near the wall. Many of the features observed in the centreplane side views of figure 9
are consistent with previous experimental investigations on similar systems (Lee et al.
1992; Ben-Yakar et al. 1998). However, contrary to the observations of Ben-Yakar &
Hanson (1998), when ignition occurs upstream or just downstream of the injection
point, flame quenching is not observed as the flow expands around the base of the
transverse jet.
For the case with the largest value of J tested in this study, we observed OH well
upstream of the injection point. Specifically, the flame is consistently anchored 5–6 jet
diameters upstream of the injection point. This feature is found only at J = 5.0 even
if burning in the shear layer is detected.
The OH layers in the shear layers are generally thin (compared with the overall
size of the jet, for example); they are convoluted and discontinuous possibly due to
local extinction. Most of the layers are organized in bulges that resemble large-scale
vortical structures of shear layers. Furthermore, moving away from the centreplane we
can observe that the reaction layers merge with regions where OH is distributed in
space and that dwell on the wall (figure 10). Connection of the shear layer with the
near-wall burning regions, which are relatively stable, might provide a mechanism for
shear layer flame stabilization.

4.2. OH layer thickness


Figure 12 shows average OH PLIF images for the J = 0.3 and J = 5.0 cases. The
average was computed over a small number of repetitions (nine for the J = 0.3 case
and 11 for the J = 5.0 case). In spite of the fact that the average is clearly not
statistically converged, it demonstrates the repeatability of the flow features shown
by the set of instantaneous images shown previously. Furthermore, it demonstrates
the intermittent nature and corrugation of the reacting shear layer that results in a
relatively broad region where the reacting shear layer typically exists. For the low-J-
value case, what is here referred to as the reacting shear layer can exist anywhere
between the wall and 3d off the wall. For the high-J-value case, the reacting shear
layer exists in a well-defined band. The thickness of this band is seen to grow from
248 M. Gamba and M. G. Mungal

E F J
A D I
1.0 B BC G
A H
C

Normalized OH profile
D
0.8 E
G
F
G
0.6 H
I
J
0.4

0.2

0
–2 –1 0 1 2

F IGURE 13. Normalized profiles of selected OH layers on the centreplane of the


J = 5.0 JISCF.

the injection point in the region of maximum deflection, and then to remain nearly
unchanged as the jet develops in the wake. In the wake, this band is approximately
3d wide.
To better quantify the properties of the reacting shear layer, some of its morphologi-
cal properties have been investigated statistically. This was done by extracting the OH
layer thickness, location and orientation from the set of instantaneous images used
to construct the average field shown in figure 12. We were able to carry out this
analysis only on the J = 5.0 case because we did not have sufficient samples for
other (lower-J) cases or the reacting shear layer was absent (for the J = 0.3 case). In
particular, the OH layers in the shear layer were extracted using an algorithm based
on the one by Tsurikov (Tsurikov 2002; Tsurikov & Clemens 2002). The algorithm
first identifies and classifies the layer, then finds its centreline, thus its direction, and
finally defines the layer profile along a line perpendicular to the local direction of
the layer. The process is shown in figure 13, where in the inset an OH PLIF image
is shown along with the positions of some profiles (labelled A–J), while the main
figure shows the (normalized) layer profiles (A–J) plotted as a function of s/λ; s is
the coordinate across the profile and λ is the width of the profile, which is defined
as the full-width at half-maximum of a Gaussian curve fitted to the profile. The solid
line shown in figure 13 indicates a normalized Gaussian demonstrating the similarity
between the central portion of the profiles and a Gaussian curve. It should be noted,
however, that the resemblance of the profiles to a Gaussian is primarily attributed to
spatial resolution limitations (blurring) introduced by the imaging system used. To
approximately account for some of these limitations, the image blurring estimation
method of Wang & Clemens (2004) was used to correct the measured layer thickness
λ by the blurring introduced by the imaging system (after its point spread function
was approximately measured).
Figure 14 shows the streamwise distribution of (a) the OH layer thickness λ
and (b) the position. In both figures, the black circles with error bars are average
values evaluated over 2d-wide vertical bands; the error bars are the corresponding
root-mean-squared (RMS) values. Furthermore, the grey circles in figure 14(b) show
the instantaneous locations of all extracted layers over the 11 frames while the solid
Reacting transverse jet in supersonic crossflow 249

(a)
0.7

(mm) 0
–5 0 5 10 15 20 25 30 35 40 45

(b) 10

0
–5 0 5 10 15 20 25 30 35 40 45

F IGURE 14. (a) Streamwise distribution of OH layer thickness, with error bars indicating
its RMS value. (b) Penetration of OH layers. Both figures refer to J = 5.0 JISCF
measurements on the centreplane. In (b) u, mean location of the OH layer with error bars
indicating its RMS value; E, instantaneous location of the OH layer extracted from the
set of instantaneous OH PLIF images; the black solid line indicates the flame boundary
extracted from the OH∗ imaging results (figure 4c).

black line is the flame boundary defined previously from the OH∗ chemiluminescence
images (see figure 4c). The figures suggest that the OH layer thickness is (on average)
approximately 0.3 mm in the near field and rapidly grows to approximately 0.5 mm.
The thickness then remains fairly constant in the wake of the JISCF.
Because the OH regions are organized in layer-like structures, one can approximate
their description as strained laminar diffusion layers (Buch & Dahm 1996, 1998;
Kothnur & Clemens 2005). In this respect, the OH layers can thus be thought of
as marking flamelets where combustion occurs, and their modelling can follow the
flamelet model (Peters 1984; Pitsch, Chen & Peters 1998). For a constant-density
steady formulation, the characteristic thickness λ of a strained laminar (non-reacting)
diffusion layer can be written as
r
D
λ∼ , (4.1)
S
where S is a characteristic (local) strain rate and D is a scalar diffusivity. Assuming
that D does not vary spatially along the shear layer of the transverse jet, the layer
thickness only inversely depends on the strain rate. We can approximate the strain
rate as the ratio between a characteristic local velocity scale and the length scale
of the transverse jet. From the (incompressible) scaling law results of Hasselbrink &
Mungal (2001), we can take the local centreline-to-free stream velocity difference and
jet penetration as suitable scale quantities in the limit that compressibility does not
play a role. Thus, for both the near and far fields, we can readily derive that the
expected dependence of λ on x/d is λ ∼ (x/d)1/2 . By this result, we would expect that
λ continuously increases with the downstream distance x/d. An increase of λ with x/d
is qualitatively observed only for the near field and the region of maximum deflection
(figure 14a); on the contrary, λ remains nearly constant for downstream distances
250 M. Gamba and M. G. Mungal
larger than x/d ≈ 20. If the approximation of the OH layers as strained diffusion layers
is correct and if their dependence on S−1/2 is accepted, then a constant λ for x/d > 20
indicates that the large-scale strain rate in the far field remains unchanged. This would
require that both the characteristic velocity and length scales remain constant or vary
at the same rate with x/d.
Figure 14(b) shows the instantaneous and ensemble-averaged positions of the OH
layers. We define this quantity as the OH flame boundary or penetration. Their mean
position is equivalent to the OH∗ flame boundary we defined previously. The OH
and OH∗ flame boundaries are thus compared in the figure (the solid line is the
OH∗ boundary). As one would expect, the OH∗ flame boundary is found to capture
the outer envelope of where the instantaneous OH layers exist. The error bars in
figure 14(b) provide a quantitative measure of the spread of the OH layer locations
around the mean value; the spread is in fact a significant fraction (approximately
1/4th) of the local penetration. It is interesting to note that unlike the subsonic
transverse jet, the trajectory defined by the OH boundary is not monotonic. In fact,
the trajectory is strongly curved close to the injection point, it contracts around
x/d ≈ 17 (as if the jet would ‘neck’), and then steadily increases in the wake.
The same contraction in the transverse jet penetration can be recognized in the
corresponding Schlieren image of figure 3(b) (near label F). From a comparison with
the Schlieren images, this feature seems to be associated with the presence of the
recompression shock (label E in figure 3b) formed on the lee side as the flow expands
around the base of the jet. A similar non-monotonic near-field penetration was also
observed in the trajectories extracted from the mixing studies of Rogers (1971b) (see
also Povinelli & Povinelli 1971); they observed a contraction of the penetration in the
region 7 < x/d < 20. It should also be pointed out that a J = 5 transverse jet is one
with a relatively low value of the momentum flux ratio (in the context of transverse
jets); thus, a near field with a well-developed jet-like behaviour is not expected
(Smith & Mungal 1998; Hasselbrink & Mungal 2001). Furthermore, in the supersonic
case there exists also a (distorted) barrel shock that might affect the evolution of
the transverse jet in the portion of maximum deflection. The combination of these
two effects could also be responsible for or at least participate in the formation of
the contraction (necking) observed at x/d ≈ 17. Because we cannot extend the same
analysis to all other cases, we cannot conclude that the necking in the OH flame
boundary is a universal feature (i.e. that it exists for the lower-J cases). However, we
note that Rogers (1971b) observed a similar feature in his non-reacting experiments
for J from 0.5 to 1.5. Finally, it is interesting to note that the point of necking of
the OH flame boundary in figure 14(b) corresponds to a local minimum in the OH
layer thickness in figure 14(a) at x/d ≈ 16. If the necking is associated with the
recompression shock, it is also possible that the recompression shock acts to further
strain the OH layers, thus decreasing their thickness locally.
To further quantify the properties of the OH layers, probability density functions
(PDFs) of the layer width λ and orientation θ (with 0 6 |θ| 6 π/2) were computed
at different downstream bands. The layer orientation θ is defined as the magnitude of
the angle between the layer and the x-axis. The resulting PDFs are shown in figure 15.
Figure 15(a) shows PDFs of the thickness λ evaluated in a region around the injector
(|x/d| < 5) and in selected downstream regions. At the injector, the distribution of λ
is relatively narrow with a peak value near 0.3 mm. At the first band downstream of
the injector (5 6 x/d 6 10), the PDF shifts to larger values and begins to broaden with
a long tail at large values of λ. Then, the distribution for x/d > 10 remains the same
at all other downstream regions; it is relatively broad with a peak value near 0.5 mm.
Reacting transverse jet in supersonic crossflow 251

(a) (b)

0.3
All
0.1
PDF

0.2

0.1

0 0.5 1.0 0 0.5 1.0 1.5


(mm) (rad)

F IGURE 15. The PDF of (a) OH layer thickness λ and (b) angular orientation |θ|.

It is also interesting to note that in all regions the PDFs show a sharp edge on the
low-λ-value side, while a longer tail on the high-value side. In general, values of less
than approximately 0.2 mm (a value that could be limited by the imaging set-up) or
larger than approximately 1 mm (a value that could be limited by our layer detection
definition and algorithm) are never detected.
Figure 15(b) shows the corresponding PDFs for the OH layer orientation angle
θ, where only its magnitude is reported. This was necessary to obtain smoother
PDFs due to the overall limited number of samples. For the same reason, only three
downstream bands were considered. The figure shows that at the injection point there
is a strong tendency of the OH layer to be nearly vertical (θ = π/2). Furthermore, it is
overall more likely to observe an orientation angle larger than π/4. On the contrary,
downstream of the injection point, the OH layers are more likely to be oriented
parallel to the free stream, θ = 0, or in any case, at an angle of less than π/4. Thus,
throughout the evolution of the windward side of the shear layer of the transverse
jet, the OH layers tend to track the interface between the jet and crossflow streams.
Furthermore, the fact that the layer orientation is mostly parallel to the free stream
in the wake indicates that the shear (i.e. velocity difference) between the wake and
the crossflow is small, or at least insufficient to generate large-scale rollers that can
promote entrainment and thus rapid mixing and combustion. In this respect, chemical
reaction along the wake of the jet progresses in a purely diffusion-dominated fashion
with minimal enhancement by large-scale turbulent structures. Thus, overall, the
wake of an isolated JISCF cannot be regarded as an efficient system for supersonic
combustion (at least for hydrogen/air combustion) consistent with wake-like behaviour.

4.3. Near-wall burning


A particular feature of interest is the distribution of OH within the boundary layer near
the surface, as was shown by figure 11(a–c). To investigate this flow feature in more
detail, plan-view OH PLIF imaging at different wall-normal locations was carried out.
Examples at selected wall-normal locations (y/d = 0.25, 0.5, 1 and >2) are shown in
figures 16–18 for the J = 0.3, 2.7 and 5.0 cases respectively. Based on the estimate of
the boundary-layer thickness at the injection point given previously and from the side-
view OH images, the measurement plane at y/d = 0.25 is well within the (undisturbed)
252 M. Gamba and M. G. Mungal

(a) 20 (b) 20
15 C1 E1 15 E1
10 10
5 5
A1 D1 D1
0 0
B1
–5 –5
–10 –10 G1
F1 F1
–15 –15
–20 –20
–10 –5 0 5 10 15 20 25 30 35 40 45 –10 –5 0 5 10 15 20 25 30 35 40 45

(c) 20 (d) 20
15 E1 15
10 10
5 5
0 D1 0
–5 –5
–10 –10
–15 –15
–20 –20
–10 –5 0 5 10 15 20 25 30 35 40 45 –10 –5 0 5 10 15 20 25 30 35 40 45

F IGURE 16. OH PLIF imaging on selected plan-view planes for the J = 0.3 JISCF. Plan
views at (a) y/d = 0.25, (b) y/d = 0.5, (c) y/d = 1 and (d) y/d = 2.

boundary layer, while that at y/d = 0.5 is approximately at the edge of it. Thus, for
each figure, panels (a–c) identify the structure of the OH regions within the boundary
layer, while panel (d) identifies the structure somewhere in the shear layer. It should
be noted that the flow features in these figures have been labelled with an alphabetical
letter followed by a number (e.g. C1), such that the letter indicates corresponding flow
features across different J cases while the number identifies the three cases reported in
the set of figures. It should be noted that labels used in figures 16–18 do not necessary
correspond to the same labels and flow features as indicated in figure 4. The main flow
and combustion features identified in figures 16–18 are also summarized in table 3.
Cases corresponding to J values lower than approximately 3 are characterized
by a plan-view imprint of the reaction zone that is consistently similar to the case
of figure 16 (J = 0.3). They are characterized by (i) long separation between the
injection hole (A1 in figure 16) and the point of ignition (C1 in figure 16), and
which is consistent with the stabilization point measurements of figure 5 extracted
from the time-averaged OH∗ imaging; (ii) a broad region where a diffuse distribution
of OH exists (E1 in figure 16); this region is several jet diameters wide and grows
with downstream distance; (iii) a characteristic spreading of the OH regions E1
downstream of the ignition point C1 that consistently grows with J (compare the
cases of figures 16 and 17); (iv) a narrow wandering region or streak (of the order
of two jet diameters in size) located along the symmetry plane where burning does
not occur (region D1 in figure 16); this non-reacting streak in the wake of the JISCF
also exists for larger J, as indicated by regions D2 and D3 in figures 17 and 18;
(v) the thickness (from the plate surface) of the region of diffuse OH is at most
Reacting transverse jet in supersonic crossflow 253

(a) 20 (b) 20
15 15
10 C2 E2 10 C2 E2
5 5
0 A2 D2 0 A2 D2

–5 B2 –5
–10 F2 G2 –10 F2
–15 –15 G2
–20 –20
–10 –5 0 5 10 15 20 25 30 35 40 45 –10 –5 0 5 10 15 20 25 30 35 40 45

(c) 20 (d) 20
15 15
10 C2 E2 10
5 5
0 D2 0
–5 –5
–10 F2 H2 –10 I2
–15 I2 –15
–20 –20
–10 –5 0 5 10 15 20 25 30 35 40 45 –10 –5 0 5 10 15 20 25 30 35 40 45

F IGURE 17. OH PLIF imaging on selected plan-view planes for the J = 2.7 JISCF. Plan
views at (a) y/d = 0.25, (b) y/d = 0.5, (c) y/d = 1.0 and (d) y/d = 3.0.

Label Identified flow feature


A Injection point
B Unburnt recirculation and horseshoe vortex regions
C Point of ignition or stabilization
D Unburnt near-wall wake streak within the boundary layer
E Regions of broad distribution of OH
F Thin OH layers in the near-field shear layer
G Finger-like OH structures penetrating the near-wall wake
H Thin OH layer in the near-wall region
I Thin OH layer in the far-field shear layer
TABLE 3. Summary of the most important flow and combustion features identified
in the near-wall burning regions as visualized by the plan-view OH PLIF images of
figures 16–18.

of the order of one jet diameter since measurements on planes above y/d = 1 only
show the reacting shear layer of the transverse jet. The thickness of the OH region
at the wall remains nearly the same for all cases, with a small initial growth with
downstream distance. This was also better shown by the side-view OH images, and
it is in agreement with the thickness of the intense OH∗ layer that was observed near
the plate surface from figure 4. In agreement with the results of figure 5, the point
of stabilization (C1 in figure 16) moves upstream and reaches the injector for J ≈ 3;
254 M. Gamba and M. G. Mungal

(a) 20 (b) 20
15 15
10 C3 E3 10 C3 E3
5 5
0 D3 0 D3
A3
–5 –5
–10 B3 F3 –10 B3 F3
–15 G3 –15 G3 I3
H3
–20 –20
–10 –5 0 5 10 15 20 25 30 35 40 45 –10 –5 0 5 10 15 20 25 30 35 40 45

(c) 20 (d) 20
15 15
10 E3 10
5 5
0 0
–5 –5
–10 F3 –10 F3
–15 I3 –15 I3
–20 –20
–10 –5 0 5 10 15 20 25 30 35 40 45 –10 –5 0 5 10 15 20 25 30 35 40 45

F IGURE 18. OH PLIF imaging on selected plan-view planes for the J = 5.0 JISCF. Plan
views at (a) y/d = 0.25, (b) y/d = 0.5, (c) y/d = 1.0 and (d) y/d = 3.0.

see point C2 on the y/d = 0.25 plane for the J = 2.7 case (figure 17). For larger J,
the stabilization point moves upstream of the injection point; see point C3 on the
y/d = 0.25 plane for the J = 5.0 case (figure 18).
The relative spanwise size of these burning regions increases with J. The broad
diffuse OH regions at the wall (y/d = 0.25) for J = 0.3 (figure 16) are of the order
of ±5d wide, and they grow to ±15d for J = 2.7 by the end of the measurement
domain (x/d = 40). The spanwise extent increases further to more than ±20d for the
J = 5.0 case. Furthermore, these regions grow laterally with an apparent origin that
approximately corresponds to the injector.
The morphological characteristics of the near-wall OH layer begin to significantly
change for the J = 5.0 case where some drastic modification takes place. The main
difference for the J = 5.0 case is that stabilization appears upstream of the injection
point – see C3 in figure 18(a) – along with the existence of a significantly large OH
footprint near the wall. We will discuss this further below, but it is believed that the
flow structure that leads to the OH imprint of figure 18 is related to the strength of the
bow shock and the importance of shock/boundary-layer interaction on the near-field
flow and mixing characteristics (Gamba et al. 2011).
Comparison of the OH imaging results for the different cases on side and planviews
(including additional side- and plan-view measurement planes available but not shown
here), suggests that the thickness of the near-wall OH regions is relatively independent
of the value of J, or at most a very weak function of it. In fact, in all cases the
edge of these burning regions is located at approximately y/d = 1. We therefore
conclude that one of the controlling flow parameters in determining the properties
Reacting transverse jet in supersonic crossflow 255
of this reacting region is the incoming boundary-layer thickness. More specifically,
we speculate that this feature is primary related to the structure and properties
of the incoming boundary layer (recall that the boundary layer is laminar and
approximately 0.5d thick in this study) and its tendency to separate upon interaction
with the bow shock. Therefore, we believe that the observed near-wall imprint is an
extensive reacting boundary layer where fuel entrainment is sustained by the upstream
recirculation region (which for sufficiently large values of J also serves as a point of
flame stabilization) and by the bow shock/jet interaction.
Imaging several diameters off the plate surface detects only the reacting shear layer
of the jet (see figures 17d–18d). Because of the small penetration of the J = 0.3 case
and the weak tendency of the shear layer to react, a well-defined reacting shear layer
is not found in this case. The reacting shear layer on y/d = 3 (J = 2.7 and 5.0) shows
characteristics similar to the OH side views shown previously; it is in fact part of the
shear layer system of the main body of the transverse jet. The lateral extent of the
jet body is significantly narrower than what is observed near the plate, being only a
few diameters wide (at most ±7d for the J = 5.0 cases, for example). By combining
information about the location of the reacting OH layers from side and plan views we
can also infer that the cross-section (i.e. on a y–z plane) shape of the reacting shear
layer has nearly a 1:1 aspect ratio.
Similarly to low-speed transverse jets, the wake of the jet is dominated by a
CVP (e.g. Santiago & Dutton 1997; Smith & Mungal 1998) generated during the
jet deflection process by reorganization of vorticity. The CVP has a significant role
in crossflow entrainment and mixing in the wake of the jet (VanLerberghe et al.
2000). For low J, the penetration of the jet is only a few jet diameters deep and
the evolution of the CVP is inevitably constrained by the presence of the wall. The
strength of the interaction of the CVP with the wall and the boundary layer would
be expected to increase with diminishing jet penetration (i.e. J). However, even the
highest-J-value case considered here remains a case with an overall low value of J
if compared with many of the transverse jets studied within the low-speed regime.
Therefore, the observed near-wall structure of the OH regions is also a result of the
constrained evolution of the CVP interaction with the near-wall flow.
Generally speaking, the OH distribution near the wall for y/d < 1 occurs over broad
contiguous regions in space. For J < 3 there are two main contiguous regions about
the streamwise plane of symmetry of the system where most of the OH is found.
Although it is difficult to establish solely from these results, this observation suggests
that the combustion process near the wall might progress either as a premixed/partially
premixed mode, or a non-premixed distributed reaction zone mode. We will return to
this below. A closer inspection of the OH fields reveals that thin OH layers are also
present at the inner interface of the jet wake. This might indicate that the combustion
process in the near-field near-wall region might proceed in a complex fashion
characterized by different but coexisting combustion modes (regimes). Examples
of thin OH layers are indicated by labels F1 for the J = 0.3 case, F2, H2 and I2
for the J = 2.7 case, and F3, H3 and I3 for the J = 5.0 case. The thin OH layers
are embedded and at times connected to the regions where OH is broadly distributed
in space and appear to be primarily associated with regions in the immediate wake
of the transverse jet. For the J = 0.3 case, features F1 and C1/E1 (see figure 17a,b)
define a two-region (or regime) burning feature: F1 is associated with localized thin
reaction at the boundary between the (fuel-rich) wake and the outer flow (effectively
this is the shear layer of the transverse jet), while E1 is the region where OH occurs
broadly. For the J = 2.7 and 5 cases, the thin OH layers exist throughout the jet.
256 M. Gamba and M. G. Mungal
For example, thin layers are found to wrap around the injector (features F2 and F3),
and to extend in the wake (H2/I2 and H3/I3) mixed with regions where OH occurs
broadly.
The set of plan-view OH images indicates that ignition in the recirculation region
in front of the transverse jet occurs only at large J (in this case only for J = 5.0)
with a characteristic bow shape, while broadly distributed OH is visible downstream
of the injector for most of the J cases. As shown in figure 18(a,b), this reacting
recirculation region wraps around the base of the jet with a characteristic ‘hole’ just
in front of the injector. Very close to the wall (y/d = 0.25, see figure 18a), the OH is
distributed uniformly and no structure is observed; on the contrary, at higher locations
(y/d = 0.5, see figure 18b) the OH is organized in stretched loops (see B3) connected
to the OH layers of the windward shear layer (F3) mixed with regions characterized
by OH streaks originating at a virtual point (for example C3 in figure 18b). The
analysis of Huber et al. (1979) on ignition and flame-holding capabilities of different
wall-injection configurations identified the upstream recirculation as a suitable location
for flame stabilization due to the long residence time and near-stagnation conditions.
Therefore, the length of the upstream recirculation region (which increases with
J) (Dowdy & Newton 1963), the low velocities associated with it (Santiago &
Dutton 1997) and the blockage effect of the jet (which increases with J) would
play an important role in this process. Another mechanism supporting ignition in the
separation region is the favourable mixing that occurs as a result of the interaction
of the issuing jet and the system of recirculation regions (Thayer 1971; Yoshida &
Tsuji 1977; Huber et al. 1979) along with the shorter ignition times expected at lean
conditions in supersonic non-premixed combustion of a cold fuel mixed with a hot
oxidizer (Huber et al. 1979) that could be observed in the upstream recirculation
region.
The burning structure in the upstream recirculation region and around the base of
the jet is not observed at J values lower than 5. However, for the other J cases, a
structure that resembles a horseshoe vortex system can still be recognized by tracking
regions of low background signal. In fact, it was found that the free stream generated
a detectable level of background signal in the images. The origin of this background
is not fully understood, but it is observed in high-temperature free streams (the
background is not present for testing conditions at lower free stream temperatures)
and not in regions where pure fuel exists. Therefore, although this background is
unwanted, it provides a convenient marker of free stream fluid and it can serve to
identify regions rich in injectant, especially in the near field and in the recirculation
region of the jet. An example of this effect is indicated in features A1, B1 and G1
in figure 16(a). Close inspection of the imaging results consistently shows a fuel-rich
region that wraps around the injector. By this method, the upstream recirculation
region and its size (for example its length) can be identified at all values of J. The
extent of this region was found to increase with J, as also shown in non-reacting
transverse jets (Dowdy & Newton 1963).
Everett et al. (1998) carried out surface oil-flow visualization around transverse
jets (at J = 1–2) to investigate the streamline pattern resulting from the interaction of
the jet with the bow shock and the boundary layer. They identified two separation
lines that wrap around the jet and that create two separate non-interacting regions.
Between these two separation lines, and between the second separation line and the
injector, two recirculation bubbles exist. The first recirculation region is associated
with the boundary layer, whereas the second recirculation region is produced by the
issuing jet (see also figure 1 for a schematic diagram). Everett et al. observed that oil
Reacting transverse jet in supersonic crossflow 257
used to mark the wall streamline did not move across these two recirculation systems.
However, from the imaging of the current study, it is clear that transfer across the
two recirculation regions has to occur to provide fuel to sustain combustion within
the large interaction region. We thus believe that the ‘hole’ visible in the reacting
upstream region is associated with the second recirculation region (which is fuel-rich),
while the reacting portion (further upstream of the ‘hole’) is the first recirculation
region where reaction can occur. The oil-flow visualization of Everett et al. also
showed that streamlines at the wall follow the trajectory of the separation streamlines
and they all wrap around the base of the jet. Furthermore, they also identified a third
recirculation region just downstream of the injector which did not interact with the
upstream separation lines. This third recirculation region originates at the injector
and extends to a triangular region downstream of the injector that is then swept
downstream in the wake of the jet. This third recirculation region defines a fuel-rich
region that might not experience sufficient free stream entrainment; thus, it could be
the origin of the unburned region developing in the wake of the jet (D3).
Many of the features observed by the oil-flow visualization experiments of Everett
et al. (1998) share many similarities with the flow structures marked by OH shown in
figure 18(a) for the J = 5 case. For example, the broad OH regions that wrap around
the jet show finger-like structures (G2 and G3 for example) in the wake of the jet that
close towards the jet centreplane in a fashion that is reminiscent of the streamlines
shown by Everett et al. The main difference between the case shown here and the
case described by Everett et al. is that the upstream recirculation region observed in
the current work does wrap around the base of the jet, but does not close behind it
as they observed instead.
One last feature of interest is the unburnt streak that develops from the injection
point into the wake near the wall (feature D in figures 16–18). This region exists
for all J cases considered, and it is very clear as the wall is approached (i.e. at
x/d = 0.25). For the J = 0.3 case (D1 in figure 16), this streak is well defined in the
wake and it is bounded by broadly diffused OH regions. For the higher-J-value cases
(see D2 and D3 in figures 17 and 18 respectively), the streak remains an isolated
feature very close to the wall (x/d = 0.25), but it becomes less distinct from the
other regions as the measurement plane is farther away from the plate. Nevertheless,
its signature can still be identified. Furthermore, at higher planes, the streak becomes
more convoluted by the reacting regions bounding them and is confined by thin
layered OH regions (I2 and I3 respectively) instead of the broader OH regions at the
lowest plane. It appears that this unburnt region might be part of, or connected to, the
fuel-rich wake of the transverse jet. In particular, it might be the part of the far-field
wake of the jet that interacts with the wall, limiting entrainment of oxidizer, and
thus preventing combustion. At low J, the wake remains within the boundary layer
and its evolution is strongly limited by the wall. For the higher-J cases, the wake
penetrates away from the boundary layer, so it can freely develop. Therefore, there
can be regions characterized by thin reaction layers away from the wall (features I2
and I3 for example), which merge with the regions of broad distributions of OH as
the wall is approached.

5. Discussion
The results of the previous sections demonstrated the strong dependence of many of
the flow features of transverse jets on J. Ignition and flame stabilization also appear
to be strongly affected by this single parameter. Near-field ignition characteristics are
258 M. Gamba and M. G. Mungal
particularly affected by it. The value of J dictates the penetration of the jet which
drives the blockage effect and induces the three-dimensional bow shock around the
jet. Interaction of the bow shock with the jet and the boundary layer triggers the
system of upstream recirculation regions. All of these processes are closely coupled
and they all affect entrainment, mixing and flame stability. The resulting complex
systems of recirculation regions seem to especially control the mixing and stability
processes. Furthermore, the dynamics of the shock system is closely coupled to the
dynamics of the underlying turbulent flow, which in turn is responsible for the mixing
between jet and crossflow fluid. This coupling has been clearly shown by the work of
Kawai & Lele (2010). On the contrary, the local properties in the wake of the jet are
less affected by J. Finally, the incoming boundary layer and how it responds to the
shock system plays a key role in the mixing (Kawai & Lele 2010) and stabilization
processes.
Generally speaking, the flame stabilizes at two different regions based on the value
of J: in the boundary layer in the wake of the transverse jet for low J or in the near-
field recirculation regions for large J. Stabilization at both locations was observed at
intermediate J.
For low J (6 1), the flame is similar to a lifted flame. Different mechanisms
are believed to contribute to the stabilization process in lifted flames (Lyons 2007;
Lawn 2009). We speculate that in this case stabilization in the wake can be reduced
to an autoignition process. For hydrogen/air chemistry, the ignition delay time tig
is limited by the H/O2 chain branching reaction and, thus, it is correlated with
temperature and pressure as tig p ∼ eΘ/T (in the context of supersonic combustion,
see for example Huber et al. 1979; Mitani, Chinzei & Kanda 2001), where Θ is a
suitable characteristic temperature. Pressure and temperature should be interpreted as
suitable local values; for example, the temperature could be interpreted as the flow
recovery temperature (Mitani et al. 2001). The pressure distribution in the near field
of the jet is highly non-uniform. Peak values of pressure depend on the value of J
(Everett et al. 1998; Gruber & Goss 1999) and they can be several times the free
stream pressure within the upstream separation region for large J and much less in the
leeward side of the near field. Further downstream, the pressure recovers to the free
stream value. Temperature may vary in a similar fashion. Thus, the time-integrated
histories of temperature and pressure could also affect the process. Comparison of tig
with a characteristic time of the flow determines whether and where local ignition
might occur.
To this end, we have calculated the ignition delay times for hydrogen/air mixtures
using the hydrogen mechanism of Hong, Davidson & Hanson (2011) for different
equivalence ratios φ of the mixture. To replicate the mixing process occurring under
the conditions of this study, we assume isobaric mixing of 1400 K air with 300 K
hydrogen in a given ratio (i.e. φ), followed by ignition. It should be noted that by
this model, the temperature of the unburned mixture varies with the equivalence ratio
– thus it will affect the ignition delay time. In practice, we have reconstructed the
model of Huber et al. (1979) with an updated chemical mechanism and conditions
specific to our study. The results at three different pressures are shown in figure 19.
The values of pressure selected here correspond to the free stream value (pa = 40 kPa),
and characteristic lower and upper values (2.5 and 4 times pa for J ∼ 1 and J ∼ 4
respectively) that are expected in the upstream recirculation region for our range of
J (estimated from measurements of similar cases, e.g. Everett et al. 1998; Gruber
& Goss 1999). The oxidizer temperature was approximated with the free stream
value. However, for regions within the boundary layer, this approximation needs
Reacting transverse jet in supersonic crossflow 259

0.3

100
0.2

50
0.1
40 kPa
100 kPa
0 160 kPa 0
0 0.5 1.0 1.5 2.0

F IGURE 19. Ignition delay time tig as a function of the equivalence ratio φ for isobaric
mixing of 1400 K air with 300 K hydrogen. The ignition point xig is approximated simply
as xig = 1/2Ua tig .

some explanation. First of all, our experiments were conducted in a high-temperature


environment, but the temperature of the flat plate remained close to the initial value
(room temperature) because of the short test times of the experiments (in fact, all
hardware remained near room temperature). This was verified by conducting wall
temperature measurements at selected locations along the expansion tube (but not on
the flat plate itself) where a temperature increase of less than 10 K over the duration
of the test was found. Furthermore, it was found that this change in wall temperature
was associated with large values of heat fluxes to the wall even if the flow evolved
in short times. The boundary layer is a compressible (Mach 2.4) boundary layer
with a constant low-temperature wall. Because of the relatively low Mach number,
viscous heating within the boundary layer is, however, expected to be small. Thus,
the maximum temperature found in the boundary layer is expected to be close to
the free stream value and not near the recovery value. This is approximately true
for a laminar low-Mach-number supersonic boundary layer over a cold wall (White
1991). Therefore, recovery temperature effects are not believed to be significant in our
configuration unless strong stagnation regions occur in the flow field, as for example
in the upstream recirculation region for large values of J.
Figure 19 shows the ignition delay time tig as a function of φ. For comparison, tig
is also converted to an equivalent ignition point xig by assuming a convective velocity
within the boundary layer equal to half the free stream value. This is just a coarse
approximation of a more complex convection process within the boundary layer and
the wake of the transverse jet. The ignition point is equivalent to the stabilization point
xs defined in figure 5. The results of figure 19 show that there is a relatively strong
nonlinear and non-monotonic dependence of the ignition delay on the mixing process
(i.e. on φ). Ignition within a reasonably short length (say, within 50d) can occur only
for a well-defined and relatively limited range of φ values: too little (too lean) or too
much (too rich) mixing may just result in flame blow-off. This suggests that there
might be a strong sensitivity of stabilization by autoignition to mixing.
The condition of minimum ignition delay time is found to be at approximately
φ = 0.25, i.e. for very lean mixtures, and not near φ = 1 as one might expect. This
is a result of competing effects on the ignition delay due to non-isothermal mixing
(φ) and temperature effects on the chemistry. It is a consequence of the fact that the
ignition delay time increases with φ → 0 (and φ → ∞), but decreases with increasing
260 M. Gamba and M. G. Mungal
temperature (and pressure). In the model constructed here, as φ is varied from φ = 0
to φ → ∞, the temperature of the unburnt mixture varies from the oxidizer value
(high temperature) to the fuel value (low temperature). Thus, lean mixtures have
higher temperatures than stoichiometric ones. The overall effect results in shorter
ignition delay times at (fuel) lean mixtures. Minimum tig at lean φ values implies
that limited mixing of fuel into the crossflow has to occur to generate a rapidly
ignitable mixture. The effect of (increasing) pressure is to lower tig for slightly
leaner mixtures. A case at higher pressure corresponds to a case with larger J where
pressure time-history effects in the near field might be present. If the initial oxidizer
temperature is increased, tig decreases significantly, but the general trend remains the
same (not shown). The case of higher oxidizer temperature also corresponds to a
case with larger J because of stagnation in the upstream recirculation region or the
presence of a stronger bow shock. Comparison of the measured point of stabilization
(figure 5) with the approximation of figure 19 is very poor, but the approximations
made have to be kept in mind. From figure 5 we observe that xs /d is approximately
20 with a weak dependence on J until J = 1 (region I in figure 5). The estimated
values are much larger than what is observed for a background pressure equal to
pa . The comparison improves if results at higher pressure are considered. Because of
the fairly significant dependence of the ignition delay on quantities associated with
the near-field processes, we conclude that the dynamics in this region maintains a
dominant role in flame stabilization even in the lifted cases (low J).
The near-wall burning zones (at y/d = 0.25) are characterized by regions where OH
occurs broadly after a relatively long lift-off that supports premixing. The observed
structure resembles the broad reaction zone of premixed or partially premixed flames
rather than the thin and localized reaction zones of non-premixed combustion, as are
found in the reacting shear layer of the jet. It should be noted, however, that the
presence of regions where OH is distributed broadly could also be a consequence
of the slow three-body recombination of OH and fast convective time (Barlow et al.
1990). Nevertheless, the observed structure of the OH regions supports our speculation
that the stabilization is autoignition-supported. However, a similar distribution of
OH is also observed for high-J cases where the flame stabilizes in the upstream
recirculation region. Finally, cases of low jet penetration confine fuel near the wall
and within the boundary layer only. Therefore, the ratio of the jet penetration to
the boundary-layer thickness should be an important parameter for the stabilization
process.
For the cases at large values of J (> 3), combustion occurs in two distinct locations:
near the wall where OH remains organized in broadly distributed regions and marks
the complex flow structure of the transverse jet, and in the shear layer with a flame
structure similar to that of a thin reaction layer. Following the previous discussion, we
further speculate that the primary stabilization mechanism at large values of J (> 3)
is associated with ignition due to shock heating and long residence times at the front
of the bow shock and within the upstream separation region. In this case, large values
of pressure and temperature exist in the near field because of strong bow shocks,
and large recirculation (separation) regions are induced as a consequence (Dowdy &
Newton 1963). They allow for very low tig compared with a time scale associated
with the residence in the separation region. Thus, the flame can stabilize directly in
the upstream separation region and in the shear layer behind the bow shock.
Figure 20 schematically summarizes the different fluid dynamic processes that are
important for the stabilization and the formation of the overall structure of reacting
transverse jets. The schematic diagram shows the 3D structure of the transverse jet
Reacting transverse jet in supersonic crossflow 261

y
S
x Bow
shock
W Jet boundary/wake

z Z Boundary layer/wake
H
Horseshoe vortex
Mach disk
Plane of U
symmetry

Upstream
Boundary recirculation
layer

Flat plate

F IGURE 20. (Colour online) Three-dimensional schematic diagram of JISCF showing the
main fuel entrainment pathways.

of figure 1 with superimposed notional streamlines (labelled A–E) to indicate the


different possible paths for entrainment of fuel into the transverse jet. A case for
large J is considered in constructing the schematic. Some of the entrainment paths
discussed here are well captured by the analysis of Kawai & Lele (2010), which
is used to formulate the various path identified here. In particular, we identify the
following.
(i) Path S: fuel ejected goes through the barrel shock system, develops into the
windward shear layer, entrains and mixes with crossflow air to form the thin
reaction layer (where the term thin is used to indicate that the OH layer is thinner
than the characteristic size of the shear layer). This path is responsible for the
formation of the strained diffusion reacting layers identified in the shear layer by
the OH PLIF imaging. Flame stabilization in this path occurs only for sufficiently
large J. The occurrence of combustion along this path is believed to be linked
to the blockage effects induced at large J.
(ii) Path W: part of the fuel ejected through the barrel shock emerges into the lee side
of the jet and streams close to the wall in the wake of the injector in a region
in between the horseshoe vortex. Fuel is continuously delivered to this region.
Combustion does not occur in this region (throughout the range of J values). This
is possibly the result of the region being too cold and fuel-rich so that ignition
cannot take place.
(iii) Path Z: fuel following path W can be transferred by turbulent transport to
the horseshoe vortex system that wraps around the base of the transverse jet.
The horseshoe vortex originates from the boundary layer, therefore an ignitable
mixture can be formed in this region. This path may be one of the two possible
paths responsible for the formation of the burning regions observed at low J
near the wall.
262 M. Gamba and M. G. Mungal
(iv) Path H: fuel leaving the injector is transferred to the upstream separation bubble
through the system of recirculation regions. This transfer process is intermittent,
highly unsteady and depends on the dynamics of the jet itself. Fuel entrained
in the separation bubble may have a residence time sufficient for ignition, thus
it is responsible for the formation of the reacting upstream separation region
seen at high-J cases. Furthermore, fuel entrained in the upstream separation
region can later be transported into the horseshoe vortex system that develops
downstream. Along with path Z, this may be the second, and possibly the
dominant, mechanism responsible for the formation of the distributed reacting
regions observed at low and intermediate J values within the boundary layer. In
other words, these reacting regions (E1, E2 and E3 in the OH PLIF plan views)
might be a signature of the horseshoe vortex system.
(v) Path U: this is another path evolving in parallel to path H. The upstream
separation bubble is a three-dimensional structure that wraps around the jet
following the bow shock. Thus, fuel entrained in the upstream separation bubble
can be transferred into the region of separated flow under the bow shock and
then ultimately washed downstream in the boundary layer. This path is relevant
only for high J (here J = 5.0) because it requires a strong and large separation
bubble with an associated strong turbulent transport (Gamba et al. 2011). This
mechanism may be responsible for the formation of regions where OH occurs
broadly distributed, as observed for J = 5.0 close to the wall (y/d 6 0.5).
As a result of the complex mixing pattern identified here, the flow field appears
to be characterized by mixed regions where chemical reaction evolves through thin
diffusion layers in the unbounded portion of the flow field, while it has features similar
to those of a distributed reaction in the region dominated by boundary-layer effects.
To some extent, this view agrees with the flame structure observed in transverse jets in
subsonic heated crossflows (Micka & Driscoll 2012). Therefore, the overall flow field
cannot be characterized simply by one combustion concept specific to one description
of the combustion process (Balakrishnan & Williams 1994; Ingenito & Bruno 2010).
Furthermore, a robust modelling approach cable of capturing the complexity of the
multiple regimes of the combustion process that may characterize these flows may be
required.

6. Conclusions
In this study we have considered reacting transverse hydrogen jets in a high-
enthalpy supersonic crossflow. We have investigated how the properties of the reacting
flow field depend on the jet-to-crossflow momentum flux ratio J. We have used OH
PLIF imaging to map the instantaneous reaction zone on several different orthogonal
planes. In particular, we have investigated the effect of J on near-wall ignition and
burning, and the role of the upstream separation region in stabilization and the
reacting shear layer.
The momentum flux ratio is found to strongly affect the ignition characteristics.
In particular, two regimes are found: (1) a low-J regime (J < 1) and (2) a high-J
regime (J > 3). The difference between the two regimes is evident from comparing
the morphological structure of the reacting flow field. For low J, the flame is lifted
and stabilizes at the wall, within the boundary layer in the wake region (in our case
at a downstream location of approximately 20–22d, weakly depending on J). The
OH occurs over broad regions in space that exist only at the wall, and appear to
track the far field of the horseshoe vortex system generated by the flow around the
Reacting transverse jet in supersonic crossflow 263
near field of the injector. The overall penetration of the transverse jet remains of the
order of the boundary-layer thickness δ. On the contrary, for the high-J cases, the
point of stabilization in the boundary layer moves at or upstream of the injector point.
The boundary layer still remains a region with intense and broadly distributed OH.
However, the shear layer is now well developed and unbounded by the boundary layer;
therefore, a well-defined reacting shear layer is sustained. The reacting shear layer
is anchored at the injector itself, perhaps stabilized by the presence of the vigorous
chemical reaction that exists upstream and around the injector, which thus serves as
the main stabilization region. It is organized in thin reacting layers that are reminiscent
of strained diffusion-dominated reacting layers, or flamelets; thus, their description
could be supported by flamelet regime arguments in spite of the large flow Reynolds
numbers (Red > 3 × 105 ). These reacting layers are remarkably similar to what is
observed in low-speed non-premixed flames (e.g. Seitzman et al. 1990; Clemens, Paul
& Mungal 1997).
Differences between the low- and high-J regimes are not only reflected by the
instantaneous flame structure and organization, but some of the global parameters
that describe the overall system also indicate the existence of a boundary at J ∼ 1
where the behaviour changes. In particular, although the penetration of the flame
(defined from the OH∗ images) does follow the classical J scaling across the J range
considered here, for J < 1 the premultiplicative factor and the exponent are strong
functions of J, while for J > 2 they tend to their limiting values (∼1 and ∼0.3
respectively). The exponent for large J compares well with that reported by previous
mixing studies. Furthermore, the integrated distribution of the heat release (inferred
by using the OH∗ signal as an approximate proxy for it) changes behaviour: for J < 1,
there is a nonlinear dependence of the slope of the far-field cumulative distribution
of the heat release and of its total (image) integrated value on J, while for J > 1 the
dependence becomes linear (as simple energy balance considerations would suggest).
The boundary between what we here define as the ‘low-’ and ‘high-’ J regimes is
probably not a universal value, but just a particular one for the specific conditions of
our experiment. In fact, it is expected that this boundary may depend on the other
working flow parameters, such as Maa , δ, Re and perhaps the molecular weight (i.e.
in variable species cases). By comparison with the flame penetration results obtained
from the OH∗ chemiluminescence imaging and qualitatively assessed from the side-
view OH PLIF images, the high/low regimes might be primarily associated with the
penetration of the transverse jet relative to the boundary layer. When the penetration
is sufficient for the transverse jet to emerge from the boundary layer, we approach the
high-J limit; on the contrary, when the jet remains in the boundary layer, it behaves
as observed in the low-J limit. In other words, the low-J limit is a case where the
overall flow is dominated by the boundary layer itself rather than by the transverse jet.
Flow blockage offered by the transverse jet to the crossflow is a fundamental
feature that sets many of the properties of the reacting flow field. Flow blockage
is directly manifested by the formation of the bow shock which, in turn, generates
the region of separated boundary layer flow upstream and around the injector itself.
The strength of the bow-shock/boundary-layer interaction controls the strength and
size of the recirculation region, and thus the morphology and strength of the reacting
boundary layer. Even though for sufficiently large J combustion is observed in both
the shear layer and the near-wall region, quantification of their relative importance
suggests that the near-wall region is the primary source of heat release (at least within
the limited domain available for the study) and it might be the dominant mechanism
for flame stabilization.
264 M. Gamba and M. G. Mungal
From the OH PLIF results, we observed that the flow field is characterized by the
coexistence of thin reaction layers and distributed reaction regions, which indicates
the coexistence of two combustion regimes. Furthermore, the two regimes interact at
times, or at least are embedded in the same region of the flow. One consequence
of this observation is that the supersonic combustion regime cannot be characterized
solely by one regime; in particular, it cannot be described solely in the limit of thin
reaction zones (Balakrishnan & Williams 1994; Ingenito & Bruno 2010). On the
contrary, a multi-regime formulation that accounts for the complexity of compressible
shear layers, boundary layers and the effects of compressibility on combustion must
be considered.

Acknowledgements
This work was sponsored by the United States National Nuclear Security
Administration (NNSA), Department of Energy, under the Predictive Science
Academic Alliance Program (PSAAP) at Stanford University, directed by P. Moin.
The authors would like to thank W. N. Heltsley and V. A. Miller for their strong
support in conducting some of the experiments presented here. The authors also
gratefully acknowledge the support of and the valuable discussion with R. K. Hanson
at Stanford University.

Appendix A
Expansion tube flow conditions are notoriously difficult to fully characterize and
control on a shot-to-shot basis. A set of facility calibration experiments was conducted
to quantify the temporal evolution of the test gas flow, its spatial uniformity and its
shot-to-shot variability. A large number of single-shot experiments were monitored,
where the primary and secondary shock speeds, along with time-history measurements
of the test gas static and pitot pressures, were measured. Using these measured
quantities, the effective bulk-averaged free stream conditions were estimated using
a zero-dimensional expansion tube model based on real gas properties constructed
based on the model of Trimpi (1962).
The primary and secondary shock speeds were measured as the times of flight
between shock counters mounted along the driven and expansion sections. Four
high-response dynamic pressure transducers (PCB 132A35) separated by a known
distance (30.5 cm) were used at each section. A signal conditioner (PCB model 481)
was used to condition the output signals of the shock counters and these were then
digitalized at 1 MHz by a 12-bit data acquisition system (National Instruments PXI
6115) interfaced and controlled by a personal computer. A high-bandwidth pressure
transducer (Kistler 603B1 conditioned by a Kistler model 5010 signal conditioner
equipped with a 180 kHz cutoff low-pass filter) positioned in the expansion section
1.2 m upstream of the test section was used to measure the static pressure of the
test gas entering the test section. The arrival of the primary shock wave at the first
shock counter of the driven section was used as the master trigger to initiate the
data acquisition, the opening of the fuel delivery valve (see below) and triggering of
any imaging equipment used in each test. Therefore, the time corresponding to the
occurrence of the master trigger is here defined as the beginning of the test, i.e. as
the time t = 0 in some of the time-history plots shown below.
The shot-to-shot variations of the shock speeds monitored over several repetitions
were found to be approximately 1.0 % and 1.7 % (one standard deviation) of the
mean values for the primary and secondary shock waves respectively. The test
Reacting transverse jet in supersonic crossflow 265

(a) Expansion gas Test gas (test time)


10
V
8 Pitot
Static
IV
6
Pitot
4 III
End of
steady flow
2 II
I Static

0
4.0 4.5 5.0 5.5

(b) Expansion gas Test gas (test time)


3

2 II IV V

I III
1

0
4.0 4.5 5.0 5.5
t (ms)

F IGURE 21. (a) Time history of the static and pitot pressure simultaneously measured
in the test section near the centreline and 65 mm downstream of the exit plane of the
expansion tube. (b) Time history of the approximate Mach number inferred from the
pitot–static measurement of (a). The schematic diagram in (a) shows the measurement
arrangement. The streamwise measurement point corresponds to the location of the fuel
injector in the JISCF experiments.

gas static pressure, pa , as measured in the expansion section was found to be


approximately 40 kPa with a shot-to-shot variation of approximately 3.2 %. Assuming
the zero-dimensional model introduced previously and assuming that the test gas
is expanded by the unsteady expansion wave to the measured test gas pressure pa ,
the nominal average temperature, speed and Mach number were then estimated to
be Ta ≈ 1400 K, Ua = 1775 m s−1 and Maa = 2.44, with corresponding shot-to-shot
variations of approximately 3 %, 2 % and 1.5 % respectively. This assessment, however,
only offers nominal bulk-averaged estimates of the aerothermal conditions of the
crossflow.
The time variations of the static and pitot pressures during a run were also
investigated. Figure 21(a) shows a typical time history of the static and pitot pressures
simultaneously measured in the test section. The time indicated on the coordinate
axis is relative to the beginning of the test. The static pressure was measured by
instrumenting a flat plate (similar to that used in the JISCF experiments) with a
flush-mounted fast-response pressure transducer (PCB 112A22), whereas the pitot
pressure was measured with a conical pitot probe where a second fast-response
pressure transducer (PCB 111A26) was recessed mounted in a cavity at the stagnation
point of the probe. A schematic diagram of the set-up is shown in figure 21(a). The
pitot probe was located at the same downstream location as the static sensor, but
266 M. Gamba and M. G. Mungal
displaced away from the wall (y-direction) by a small distance (1.2 probe body
diameters) in a position that would not interfere with the static port. The static sensor
was located on the flat plate at the same relative location where the jet injector was
located in the combustion experiments (approximately 65 mm from the exit of the
expansion tube). For the static port, the active area of the transducer is approximately
4 mm in diameter, which limits its spatial resolution and response. The measured
pitot–static pressure of figure 21(a) was then converted to a time history of the local
Mach number in a neighbouring region of the measurement points using the Rayleigh
pitot equation. In converting to the Mach number, a ratio of specific heats, γ , of 5/3
was assumed for the helium flow portion (region II), whereas a value of γ = 1.31
was used for the air portion (III–IV), as estimated from the NASA polynomials at
a temperature Ta . Clearly, this arrangement does not provide a truly instantaneous
and coinciding measurement of the quantities of interest, and some error would
be expected (although not quantified). Nevertheless, it was used to investigate the
temporal evolution of the flow characteristics.
Figure 21(a) shows a representative example of the time history of the static and
pitot pressures during a typical run (normalized by pa ). Referring to the gas dynamic
process occurring during operation of the expansion tube (Trimpi 1962) and to the
labels indicated in figure 21(a), upon arrival of the secondary shock (point I, at
approximately 4.1 ms) travelling within the expansion gas, the static pressure jumps
from the quiescent value to the post-shock value (near pa ). However, because of the
pressure mismatch between the incoming expansion gas flow exiting the tube and the
nearly constant ambient pressure in the test section (maintained by the dump tank), the
expansion gas undergoes further (isentropic) expansion as it leaves the expansion tube.
Therefore, after the shock jump, the static pressure drops (and the flow accelerates)
in region II. Similarly, the pitot pressure jumps at the arrival of the secondary shock
wave, and it remains constant during the expansion gas flow. The slow rise of the
pitot pressure at the arrival of the shock is a result of the design of the pitot probe
which attenuates the temporal response of the probe. After the expansion gas outflow,
the expansion/test gas contact surface arrives at approximately 4.6 ms and the test gas
(air) follows, as the increase in pitot pressure indicates. The static pressure also rises
to the test gas value (near pa = 40 kPa). Across the contact surface, static pressure
and speed should be preserved, whereas the temperature and fluid (hence Mach
number) vary. It should be noted that, in general, the contact surface might not truly
be a well-defined material discontinuity, but, due to mixing, it might be a distributed
region of finite size. Therefore, the slow rise in static pressure is a consequence of
the finite-size nature of the contact surface combined with the transient onset of the
test gas flow structure at the exit of the expansion tube. Similarly, the slow rise in
pitot pressure results from the combined effect of contact surface spreading, the flow
establishment process and the finite temporal response of the probe. During region
IV, which extends from 4.75 to 5.25 ms, both the static and the pitot pressures
remain nearly constant. The arrival of the disruptive expansion wave near 5.25 ms
(V) results in an increase in pitot pressure and, therefore, concludes the stationary
test time, and hence the temporal window where experiments can be conducted under
nominally stationary conditions. Typically, the test time duration (region IV) for the
flow conditions considered in the study is found to be approximately 500 µs. The
single-shot imaging experiments presented here were all taken during the late portion
of the test time, region IV.
The simultaneous measure of the pitot and static pressure is then converted to the
corresponding flow Mach number and it is shown in the time history of figure 21(b).
Reacting transverse jet in supersonic crossflow 267

1.2

1.0

0.8

0.6

0.4
Injector
0.2

0
0 10 20 30 40 50

F IGURE 22. Test time-averaged static pressure distribution, hpw i/pa , along the flat plate
centreline. The error bars indicate the pressure standard deviation computed during the test
time.

The supersonic value of the Mach number during the expansion gas flow in region II
is a result of the further expansion the flow undergoes while leaving the expansion
tube (the nominal post-shock flow Mach number of the expansion gas is only near
unity). After the expansion gas flow, the test gas Mach number climbs to the test time
value, near Maa = 2.43. To confirm the measured flow Mach number Maa , another
independent set of calibration runs was performed where the average Maa was inferred
from a measure of the oblique shock angle formed about a 20◦ ramp placed in the
test section. With this approach, the estimated Maa , based on nine different repetitions
under nominally identical conditions, was found to be Maa = 2.4 ± 2 %, which is in
agreement with the pitot–static measurement of figure 21(b). All of the calibration
measurements introduced to determine the bulk-averaged free stream conditions of the
experiment indicate that they are near Maa = 2.4, pa = 40 kPa and Ta ≈ 1400 K, as
indicated previously.
The results of figure 21 indicate that the static and pitot pressures and Mach number
are not truly constant during the test time (region IV). The pitot pressure, for example,
shows a low-frequency variation over the test time; on the contrary, the static pressure
shows a systematic pressure increase (up to 5–10 % of the average value) during the
late portion of the test time. Inspection of several repetitions indicates that during
the test time the measured value of these quantities is within ±5 % (defined as one
standard deviation) of their corresponding average value during the test time. To aid
the reader, in figure 21, the average value (during the test time) for each quantity
is shown as the dot-dashed line and the ±5 % bounds around the average value are
shown by the dashed lines.
Time histories of the wall static pressure of the type of figure 21(a) were then
used to infer the average pressure distribution along the flat plate centreline (hpw i),
as shown in figure 22. This distribution was constructed by carrying out wall static
pressure measurements with an arrangement similar to that schematically shown
in figure 21(a), but with the pitot probe removed and two other static pressure
transducers mounted on the flat plate spaced 19 mm and 38 mm from the first one.
Then, the instrumented flat plate was positioned at different downstream locations to
map the axial distribution of the pressure. As figure 22 shows, hpw i remains constant
in the vicinity of the injector, there is approximately an 8 % increase towards the
end of the plate and then the pressure rolls off as a result of the end of the region
268 M. Gamba and M. G. Mungal

(a)
15
Test time

10
Test gas
Secondary arrival
shock arrival
5

Test time
termination
0
4.0 4.5 5.0 5.5
t (ms)

(b)
10 –0.7 0.7

0
–1.0 –0.5 0 0.5 1.0

F IGURE 23. (a) Time history of the pitot pressure, pt /pa , simultaneously measured at
three radial locations on the expansion tube exit plane. (b) Radial profile of the test time
average pitot pressure, hpt i/pa , measured at the exit plane over several runs; the error bar
indicates the standard deviation computed during the test time.

of uniform flow exiting the expansion tube. In the experiments presented here, the
region of interest where measurements were performed extended from upstream of the
injector (x/d = 0) to approximately x/d = 40 where the region of uniform flow ends.
To conclude this section investigating the flow properties of the crossflow, the radial
profile of the pitot pressure pt was reconstructed by scanning a pitot probe similar
to that described in the schematic diagram of figure 21(a). The pt profile was used
as an assessment of the radial uniformity of the flow relative to the size of the flat
plate used in the experiment. The pitot rack that was described in the work of Örley
et al. (2011) was utilized. The rack was constructed with five equally spaced pitot
probes and it could be positioned at five different radial locations, effectively providing
25 measurement points across the full diameter of the expansion tube over different
repetitions. A more detailed description of the pitot probe rack can be found in Örley
et al. (2011).
Figure 23(a) shows the time history of pt simultaneously measured at three radial
locations (r/R = 0, 0.4 and 0.8, where R = 7 cm is the tube radius). The same general
characteristics as described in figure 21 can be observed in the results of this figure.
During the test time, the central portion of the flow shows a similar evolution of pt ,
indicating radial uniformity of the properties, whereas close to the wall the growing
boundary layers affect the evolution of pt . The radial profile of the pitot pressure
Reacting transverse jet in supersonic crossflow 269

1400
Test time
1200
1000 He gas
arrival
800 Air
arrival
600
400
200
0
1 2 3 4 5 6
t (ms)

F IGURE 24. Example of the average time history of the plenum pressure, pj,o , measured
over multiple repetitions under identical fuelling conditions. The error bar refers to three
times the standard deviation computed over multiple repetitions. During the indicated test
time, the plenum pressure for this representative case is maintained constant to within
3 kPa.

averaged over the test time, hpt i, is shown in figure 23(b) and it was constructed from
plots of the type shown in figure 23(a) over several repetitions while the pitot rack
was scanned over the diameter. At least two runs were performed at each position of
the rack. The radial profile indicates that over the central portion within |r/R| < 0.7
the flow is, on average, uniform to within 4 % of the bulk-averaged value and only the
region |r/R| > 0.7 is affected by the boundary layer. Therefore, the flat plate (which
is approximately 1.4R wide) utilized in the experiments reported here is subject to a
radially uniform incoming flow.

Appendix B
The injector was calibrated by measuring the plenum pressure, pj,o , and assuming
isentropic expansion to sonic discharge conditions at the exit plane. Pressure
measurements were performed with a fast pressure transducer (PCB 113B26,
conditioned and digitized in a similar way to the pressure sensors on the expansion
tube) recess mounted in the plenum of the injector through a 2 mm long, 1 mm
diameter orifice. The transducer signal was low-pass filtered at 80 kHz by an
eight-pole Butterworth filter before being digitized by the data acquisition system.
Measurements were performed in a static free stream and in the presence of crossflow;
both conditions gave similar temporal profiles of pj,o . A representative example of
the time history of pj,o is shown in figure 24. The figure shows the average time
history measured over several repetitions under identical conditions (in static ambient
at pa = 40 kPa). The error bars indicate three times the standard deviation computed
at each point over the set of repetitions (only a few selected points are shown for
clarity). Following triggering of the solenoid valve at t = 0 (which corresponds to
the beginning of the test), a slow increase of the plenum pressure is observed until
a steady value is reached at approximately 3.5 ms. For the flow conditions selected
in the experiment, it is found (see appendix A) that the expansion gas arrives at
approximately 4.25 ms and the test gas follows at approximately 4.75 ms. During the
test time (4.75–5.25 ms), the plenum pressure typically remains constant to within
0.5 % of the indicated average value.
270 M. Gamba and M. G. Mungal
REFERENCES

A LLEN , M. G., PARKER , T. E., R EINEEKE , W. G., L EGNER , H. H., F OUTTER , R. R., R AWLINS ,
W. T. & DAVIS , S. J. 1993 Fluorescence imaging of OH and NO in a model supersonic
combustor. AIAA J. 31, 3, 505–512.
AYOOLA , B. O., B ALACHANDRAN , R., F RANK , J. H., M ASTORAKOS , E. & K AMINSKI , C. F.
2006 Spatially resolved heat release rate measurements in turbulent premixed flames. Combust.
Flame 144, 1–16.
B ALAKRISHNAN , G. & W ILLIAMS , F. A. 1994 Turbulent combustion regimes for hypersonic
propulsion employing hydrogen–air diffusion flames. J. Propul. Power 10, 3, 434–437.
BARLOW, R. S., D IBBLE , R. W., C HEN , J.-Y. & L UCHT, R. P. 1990 Effect of Damköhler number
on superequilibrium OH concentration in turbulent nonpremixed jet flames. Combust. Flame
82, 235–251.
B EN -YAKAR , A. & H ANSON , R. K. 1998 Experimental investigation of flame-holding capability of
hydrogen transverse jet in supersonic cross-flow. Proc. Combust. Inst. 27, 2173–2180.
B EN -YAKAR , A., K AMEL , M., M ORRIS , C. & H ANSON , R. K. 1998 Hypersonic combustion and
mixing studies using simultaneous OH PLIF and Schlieren imaging. In 36th AIAA Aerospace
Sciences Meeting and Exhibit, Reno, NV.
B EN -YAKAR , A., M UNGAL , M. G. & H ANSON , R. K. 2006 Time evolution and mixing characteristics
of hydrogen and ethylene transverse jets in supersonic crossflows. Phys. Fluids 18, 026101.
B IER , K., K APPLER , G. & W ILHELMI , H. 1971 Influence of the injection conditions on the ignition
of methane and hydrogen in a hot Mach 2 air stream. AIAA J. 9, 9, 1865–1866.
B ILLIG , F. S., O RTH , R. C. & L ASKY, M. 1971 A unified analysis of gaseous jet penetration.
AIAA J. 9, 6, 1048–1058.
B UCH , K. A. & DAHM , W. J. A. 1996 Experimental study of the fine scale structure of the conserved
scalar mixing in turbulent shear flow. I Sc > 1. J. Fluid Mech. 317, 21–71.
B UCH , K. A. & DAHM , W. J. A. 1998 Experimental study of the fine scale structure of the conserved
scalar mixing in turbulent shear flow. II Sc ≈ 1. J. Fluid Mech. 364, 1–29.
C LEMENS , N. T. 2002 Encyclopedia of imaging science and technology. In Flow Imaging,
pp. 390–419. Wiley.
C LEMENS , N. T., PAUL , P. H. & M UNGAL , M. G. 1997 The structure of OH fields in high Reynolds
number turbulent jet diffusion flames. Combust. Sci. Technol. 129, 1, 165–184.
C OHEN , L. S., C OULTER , L. J. & E GAN , W. J. 1971 Penetration and mixing of multiple gas jets
subject to a crossflow. AIAA J. 9, 4, 718–724.
C ORTELEZZI , L. & K ARAGOZIAN , A. R. 2001 On the formation of the counter-rotating vortex pair
in transverse jets. J. Fluid Mech. 446, 347–373.
C RIST, S., S HERMAN , P. M. & G LASS , D. R. 1966 Study of the highly underexpanded sonic jet.
AIAA J. 4, 1, 68–71.
D ONBAR , J. M., G RUBER , M. R., JACKSON , T. A., C ARTER , C. D. & M ATHUR , T. 2000 OH
planar laser-induced fluorescence imaging of hydrocarbon-fueled scramjet combustor. Proc.
Combust. Inst. 28, 1, 679–687.
D OSTER , J. C., K ING , P. I., G RUBER , M. R., C ARTER , C. D., RYAN , M. D. & H SU , K.-Y. 2009
In-stream hypermixer fueling pylons in supersonic flow. J. Propul. Power 25, 4, 885–901.
D OWDY, M. W. & N EWTON , J. F. 1963 Investigation of liquid and gaseous secondary injection
phenomena on a flat plate with M = 2.01 to M = 4.54. JPL Tech. Rep. No. 32-542.
E VERETT, D. E., W OODMANSEE , M. A., D UTTON , J. C. & M ORRIS , M. J. 1998 Wall pressure
measurements for a sonic jet injected transversely into a supersonic crossflow. J. Propul. Power
14, 6, 861–868.
F ERRI , A. 1973 Mixing-controlled supersonic combustion. Annu. Rev. Fluid Mech. 5, 301–338.
F RIC , T. F. & ROSHKO , A. 1994 Vortical structure in the wake of a transverse jet. J. Fluid Mech.
279, 1–47.
G AMBA , M., T ERRAPON , V. E., S AGHAFIAN , A., M UNGAL , M. G. & P ITSCH , H. 2011 Assessment
of the combustion characteristics of hydrogen transverse jets in supersonic crossflow. In Annual
Research Briefs, Center for Turbulence Research, Stanford University, pp. 259–272.
Reacting transverse jet in supersonic crossflow 271
G AUBA , G., K LAVUHN , K. G., M C DANIEL , J. C., V ICTOR , K. G., K RAUSS , R. H. &
W HITEHURST III, R. B. 1997 OH planar laser-induced fluorescence velocity measurements
in a supersonic combustor. AIAA J. 35, 4, 678–686.
G ÉNIN , F. & M ENON , S. 2010 Dynamics of sonic jet injection into supersonic crossflow. J. Turbul.
11, 4, 1–30.
G ROUT, R., G RUBER , A., Y OO , C. & C HEN , J. 2011 Direct numerical simulation of flame stabilization
downstream of a transverse fuel jet in cross-flow. Proc. Combust. Inst. 33, 1, 1629–1637.
G RUBER , M. R. & G OSS , L. P. 1999 Surface pressure measurements in supersonic transverse injection
flowfields. J. Propul. Power 15, 5, 633–641.
G RUBER , M. R., N EJAD , A. S., C HEN , T. H. & D UTTON , J. C. 1995 Mixing and penetration
studies of sonic jets in a Mach 2 freestream. J. Propul. Power 11, 2, 315–323.
G RUBER , M. R., N EJAD , A. S., C HEN , T. H. & D UTTON , J. C. 1996 Bow shock/jet interaction in
compressible transverse injection flowfields. AIAA J. 34, 10, 2191–2193.
G RUBER , M. R., N EJAD , A. S., C HEN , T. H. & D UTTON , J. C. 1997a Compressibility effects in
supersonic transverse injection flowfields. Phys. Fluids 9, 5, 1448–1461.
G RUBER , M. R., N EJAD , A. S., C HEN , T. H. & D UTTON , J. C. 1997b Large structure convection
velocity measurements in compressible transverse injection flowfields. Exp. Fluids 22, 397–407.
G UTMARK , E. J., S CHADOW, K. C. & Y U , K. H. 1995 Mixing enhancement in supersonic free
shear flows. Annu. Rev. Fluid Mech. 27, 375–417.
H ARTFIELD , R. J. & B AYLEY, D. J. 1995 Experimental investigation of angled injection in a
compressible flow. J. Propul. 12, 2, 442–445.
H ARTFIELD , R. J., H OLLO , S. D. & M C DANIEL , J. C. 1994 Experimental investigation of a
supersonic swept ramp injector using laser-induced iodine fluorescence. J. Propul. Power
10, 1, 129–135.
H ASSELBRINK , E. F. & M UNGAL , M. G. 2001 Transverse jets and jet flames. Part 1. Scaling laws
for strong transverse jets. J. Fluid Mech. 443, 1–25.
H ELTSLEY, W. N., S NYDER , J. A., C HEUNG , C. C., M UNGAL , M. G. & H ANSON , R. K.
2007 Combustion stability regimes of hydrogen jets in supersonic crossflow. In 43rd
AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, Cincinnati, OH, July 8–11,
AIAA-2007-5401.
H ELTSLEY, W. N., S NYDER , J. A., H OULE , A. J., DAVIDSON , D., M UNGAL , M. G. & H ANSON ,
R. K. 2006 Design and characterization of the Stanford 6 inch expansion tube. In 42nd
AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, 9–12 July, AIAA-2006-4443.
H ERSCH , M., P OVINELLI , L. A. & P OVINELLI , F. P. 1970 Optical study of sonic and supersonic
jet penetration from a flat plate into a Mach 2 airstream. NASA Tech. Rep. TN D-5717.
H OLLO , S. D., M C DANIEL , J. C. & H ARTFIELD , R. J. 1994 Quantitative investigation of compressible
mixing: staged transverse injection into Mach 2 flow. AIAA J. 32, 3, 528–534.
H ONG , Z., DAVIDSON , D. & H ANSON , R. 2011 An improved H2/O2 mechanism based on recent
shock tube/laser absorption measurements. Combust. Flame 158, 633–644.
H SU , K.-Y., C ARTER , C. D., G RUBER , M. R., B ARHORST, T. & S MITH , S. 2010 Experimental
study of cavity-strut combustion in supersonic flow. J. Propul. Power 26, 6, 1237–1246.
H UANG , R. & C HANG , J. 1994 The stability and visualized flame and flow structures of a combusting
jet in cross flow. Combust. Flame 98, 3, 267–278.
H UBER , P. W., S CHEXNAYDER , C. J. & M C C LINTON , C. R. 1979 Criteria for self-ignition of
supersonic hydrogen–air mixtures. NASA Tech. Rep. TP-1457.
I NGENITO , A. & B RUNO , C. 2010 Physics and regimes of supersonic combustion. AIAA J. 48, 3,
515–525.
K ARAGOZIAN , A. R. 2010 Transverse jets and their control. Prog. Energy Combust. Sci. 36, 5,
531–553.
K AWAI , S. & L ELE , S. K. 2010 Large-eddy simulation of jet mixing in supersonic crossflows. AIAA J.
48, 9, 2063–2083.
K ELSO , R. M., L IM , T. T. & P ERRY, A. E. 1996 An experimental study of round jets in cross-flow.
J. Fluid Mech. 306, 111–144.
272 M. Gamba and M. G. Mungal
K ELSO , R. M. & S MITS , A. J. 1995 Horseshoe vortex systems resulting from the interaction between
a laminar boundary layer and a transverse jet. Phys. Fluids 7, 153–158.
K OBAYASHI , K., B OWERSOX , R. D. W., S RINIVASAN , R., C ARTER , C. D. & H SU , K.-Y. 2007
Flowfield studies of diamond-shaped fuel injector in a supersonic flow. J. Propul. Power 23,
6, 1168–1176.
K OLLA , H., G ROUT, R. W., G RUBER , A. & C HEN , J. H. 2012 Mechanisms of flame stabilization and
blowout in a reacting turbulent hydrogen jet in cross-flow. Combust. Flame 159, 8, 2755–2766.
K OTHNUR , P. S. & C LEMENS , N. T. 2005 Effects of unsteady strain rate on scalar dissipation
structures in turbulent planar jets. Phys. Fluids 17, 125105.
L AWN , C. J. 2009 Lifted flames on fuel jets in co-flowing air. Prog. Energy Combust. Sci. 35, 1–30.
L EE , J. G. & S ANTAVICCA , D. A. 2003 Experimental diagnostics for the study of combustion
instabilities in lean premixed combustors. J. Propul. Power 19, 5, 735–750.
L EE , M. P., M C M ILLIN , B. K., PALMER , J. L. & H ANSON , R. K. 1992 Planar fluorescence imaging
of a transverse jet in a supersonic crossflow. AIAA J. 8, 4, 729–735.
L IN , K.-C., RYAN , M., C ARTER , C., G RUBER , M. & R AFFOUL , C. 2010 Raman scattering
measurements of gaseous ethylene jets in Mach 2 supersonic crossflow. J. Propul. Power 26,
3, 503–513.
LYONS , K. M. 2007 Toward an understanding of the stabilization mechanisms of lifted turbulent jet
flames: experiments. Prog. Energy Combust. Sci. 33, 211–231.
M AHESH , K. 2013 The interaction of jets with crossflow. Annu. Rev. Fluid Mech. 45, 379–407.
M ARGASON , R. J. 1993 Fifty years of jet in crossflow research. In AGARD Conference Proceeding.
M C C LINTON , C. R. 1974 Effect of ratio of wall boundary layer thickness to jet diameter on mixing
of a normal hydrogen jet in a supersonic stream. NASA TM X-3030.
M C DANIEL , J. C. & G RAVES , J. 1988 Laser-induced-fluorescence visualization of transverse gaseous
injection in a nonreaction supersonic combustor. J. Propul. 4, 6, 591–597.
M C M ILLIN , B. K., S EITZMAN , J. M. & H ANSON , R. K. 1994 Comparison of NO and OH
planar fluorescence temperature measurements in scramjet model flowfields. AIAA J. 32, 10,
1945–1952.
M ICKA , D. J. & D RISCOLL , J. F. 2012 Stratified jet flames in a heated (1390k) air cross-flow with
autoignition. Combust. Flame 159, 3, 1205–1214.
M ITANI , T., C HINZEI , N. & K ANDA , T. 2001 Reaction and mixing-controlled combustion in scramjet
engines. J. Propul. Power 17, 308–314.
Ö RLEY, F., G AMBA , M., A DAMS , N. A. & I ACCARINO , G. 2012 A study of expansion tube gas flow
conditions for scramjet combustor model testing. In 42nd AIAA Fluid Dynamics Conference
and Exhibit, AIAA-2012-3264.
Ö RLEY, F., S TRAND , C. L., M ILLER , V. A., G AMBA , M. & I ACCARINO , G. 2011 A study of
expansion tube gas flow conditions for scramjet combustor model testing. In Annual Research
Briefs, Center for Turbulence Research, Stanford University, pp. 285–296.
PAPAMOSCHOU , D. & H UBBARD , D. G. 1993 Visual observations of supersonic transverse jets. Exp.
Fluids 14, 468–476.
P ETERS , N. 1984 Laminar diffusion flamelet models in non-premixed turbulent combustion. Prog.
Energy Combust. Sci. 10, 319–329.
P ETERSON , D. M. & C ANDLER , G. V. 2010 Hybrid Reynolds-averaged and large-eddy simulation
of normal injection into a supersonic crossflow. J. Propul. Power 26, 3, 533–544.
P ITSCH , H., C HEN , M. & P ETERS , N. 1998 Unsteady flamelet modeling of turbulent hydrogen–air
diffusion flames. Proc. Combust. Inst. 27, 1057–1064.
P ORTZ , R. & S EGAL , C. 2006 Penetration of gaseous jets in supersonic flows. AIAA J. 44, 10,
2426–2429.
P OVINELLI , F. P. & P OVINELLI , L. A. 1971 Correlation of secondary sonic and supersonic gaseous
jet penetration into supersonic crossflows. NASA Tech. Rep. TN D-6370.
R ASMUSSEN , C. C., D HANUKA , S. K. & D RISCOLL , J. F. 2007 Visualization of flameholding
mechanisms in a supersonic combustor using PLIF. Proc. Combust. Inst. 31, 2, 2505–2512.
Reacting transverse jet in supersonic crossflow 273
R ASMUSSEN , C. C., D RISCOLL , J. F., H SU , K.-Y., D ONBAR , J. M., G RUBER , M. R. & C ARTER ,
C. D. 2005 Stability limits of cavity-stabilized flames in supersonic flow. Proc. Combust. Inst.
30, 2825–2833.
ROGERS , R. C. 1971a Mixing of hydrogen injected from multiple injectors normal to a supersonic
airstream. NASA Tech. Rep. TN D-6476.
ROGERS , R. C. 1971b A study of the mixing of hydrogen injected normal to a supersonic airstream.
NASA Tech. Rep. TN D-6114.
ROTHSTEIN , A. D. 1992 A study of the normal injection of hydrogen into a heated supersonic flow
using planar laser-induced fluorescence. PhD thesis, Massachusetts Institute of Technology.
ROTHSTEIN , A. D. & WANTUCK , P. J. 1992 A study of normal injection of hydrogen into a heated
supersonic flow using planar laser-induced fluorescence. In 28th Joint Propulsion Conference
and Exhibit, AIAA-1992-3423.
RYAN , M., G RUBER , M. R., C ARTER , C. D. & M ATHUR , T. 2009 Planar laser-induced fluorescence
imaging of OH in a supersonic combustor fueled with ethylene and methane. Proc. Combust.
Inst. 32, 2, 2429–2436.
S ANTIAGO , J. G. & D UTTON , J. C. 1997 Velocity measurements of a jet injected into a supersonic
crossflow. J. Propul. Power 13, 2, 264–273.
S CHAUPP, C., F RIEDRICH , R. & F OYSI , H. 2012 Transverse injection of a plane-reacting jet into
compressible turbulent channel flow. J. Turbul. 13, no. 24.
S CHETZ , J. A. & B ILLIG , F. S. 1966 Penetration of gaseous jets injected into a supersonic stream.
J. Spacecr. 3, 11, 1658–1665.
S CHETZ , J. A., H AWKINS , P. F. & L EHMAN , H. 1967 Structure of highly underexpanded transverse
jets in a supersonic stream. AIAA J. 5, 5, 882–884.
S EINER , J. M., DASH , S. M. & K ENZAKOWSKI , D. C. 2001 Historical survey on enhanced mixing
in scramjet engines. J. Propul. Power 17, 6, 1273–1286.
S EITZMAN , J. M., U NGUT, A., PAUL , P. H. & H ANSON , R. K. 1990 Imaging and characterization
of OH structures in a nonpremixed turbulent flame. Proc. Combust. Inst. 23, 637–644.
S MITH , S. H. & M UNGAL , M. G. 1998 Mixing, structure and scaling of the jet in crossflow. J. Fluid
Mech. 357, 83–122.
S U , L. & M UNGAL , M. G. 2004 Simultaneous measurements of scalar and velocity field evolution
in turbulent crossflowing jets. J. Fluid Mech. 513, 1–45.
S ULLIVAN , R., W ILDE , B., N OBLE , D. R., S EITZMAN , J. M. & L IEUWEN , T. C. 2014 Time-averaged
characteristics of a reacting fuel jet in vitiated cross-flow. Combust. Flame 161, 7, 1792–1803.
T HAYER , W. J. 1971 The two-dimensional separated flow region upstream of inert and chemically
reactive transverse jets. Tech. Rep. D1-82-1066. Flight Sci. Lab., Boeing Sci. Res. Lab.
T ORRENCE , M. G. 1971 Effect of injectant molecular weight on mixing of a normal jet in a Mach
4 airstream. NASA Tech. Rep. TN D-6061.
T RIMPI , R. L. 1962 A preliminary theoretical study of the expansion tube, a new device for producing
high-enthalpy short-duration hypersonic gas flows. NASA Tech. Rep. TR R-133.
T SURIKOV, M. S. 2002 Experimental investigation of the fine scale structure in turbulent gas-phase
jet flows. PhD thesis, The University of Texas.
T SURIKOV, M. S. & C LEMENS , N. T. 2002 The structure of dissipative scales in axisymmetric
turbulent gas-phase jets. In AIAA Aerospace Science Meeting, AIAA-2002-0164.
VAN L ERBERGHE , W. M., S ANTIAGO , J. G., D UTTON , J. C. & L UCHT, R. P. 2000 Mixing of a
sonic transverse jet injected into a supersonic flow. AIAA J. 38, 3, 470–479.
V ERGINE , F., M ADDALENA , L., M ILLER , V. & G AMBA , M. 2015 Supersonic combustion of pylon
injected hydrogen in high-enthalpy flow with imposed vortex dynamics. J. Propul. Power 31,
1, 89–103.
WANG , G. & C LEMENS , N. T. 2004 Effects of imaging system blur on measurements of flow scalars
and scalar gradients. Exp. Fluids 37, 194–205.
W HITE , F. M. 1991 Viscous Fluid Flow. McGraw-Hill.
Y OSHIDA , A. & T SUJI , H. 1977 Supersonic combustion of hydrogen in vitiated airstream using
transverse injection. AIAA J. 15, 4, 463–464.

You might also like