You are on page 1of 13

Combustion and Flame 193 (2018) 440–452

Contents lists available at ScienceDirect

Combustion and Flame


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

Local extinction mechanisms analysis of spray jet flame using high


speed diagnostics
Antoine Verdier, Javier Marrero Santiago, Alexis Vandel, Gilles Godard, Gilles Cabot,
Bruno Renou∗
INSA Rouen, UNIROUEN, CNRS, CORIA, Normandie Universite, Rouen 76000, France

a r t i c l e i n f o a b s t r a c t

Article history: This paper reports an experimental study where flame structure, flow topology and local extinction mech-
Received 14 October 2017 anisms of n-heptane spray flames are investigated. The burner consists of an annular non-swirling co-flow
Revised 23 March 2018
of air that surrounds a central hollow-cone spray injector, leading to a lifted spray flame. The experiments
Accepted 23 March 2018
include measurements of droplet size and velocity by Phase Doppler Anemometry (PDA), flame structure
by High-Speed Planar Laser Induced Fluorescence of OH radical (HS-OH-PLIF) simultaneously recorded
Keywords: with the velocity fields of the reactive flow obtained by High-Speed Particle Image Velocimetry (HS-PIV).
Spray jet flame The poly-disperse spray distribution yields small droplets along the centerline axis while the majority of
Flame structure the mass is located as large droplets along the spray borders. These large droplets associated with high
High-speed OH-PLIF and PIV
velocities have ballistic trajectories and strongly interact with the inner wrinkled partially premixed flame
Flame extinction
front and the outer diffusion flame front. Simultaneous HS-OH-PLIF and HS-PIV images characterize the
dynamics of extinction events in the spray jet flame. In the inner reaction zone, local flame extinctions
are mainly controlled by the shear layer induced by the co-flow and the fuel–air heterogeneities due to
the evaporation of small droplets in the vicinity of the flame front. The large scales of turbulence in the
shear layer play a significant role in the dynamics of these extinctions. It is also found that the large
inertial droplets penetrate the lower part of the inner front reaching the burned gases, where they evap-
orate rapidly. They also disturb the outer reaction zone due to the low droplets temperature and the rich
mixture in the wake of droplet. These new results on local extinction of spray flames and droplet–flame
interactions will also strengthen the CORIA Rouen Spray Burner (CRSB) database for the improvement of
evaporation and combustion models for reacting sprays.
© 2018 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

1. Introduction [1–3] investigated the flame structures in the stabilization zones


of a two-phase jet flame above a coaxial air-blast injector. From
New burner designs must move onto configurations with lower quantitative OH Planar Laser Induced Fluorescence (OH-PLIF) mea-
pollutant emissions and higher efficiency. This requires a funda- surements, authors suggested that the flame presented a struc-
mental understanding of the processes involved in two-phase and ture of two diverging and opposed diffusion-like fronts. In their
lean combustion. Two-phase combustion carries coupled multi- Jet Spray Flame burner, operating with a pressurized fuel injec-
physical and chemical phenomena which generally take place si- tor and a large bluff-body, Friedmann et al. [4] showed that the
multaneously within the combustion chambers. To overcome the spray flame also may have a dual reaction zone structure consist-
difficulties resulting from real geometries, the jet spray flame ing of an outer diffusion flame and an inner partially-premixed
configuration is taken as a reference for many experimental in- flame. However, changing either the co-flow rate [5,6] or the co-
vestigations. This canonical configuration includes a central pres- flow composition [7] affects the flame topology, and local extinc-
sure or air-blast fuel injector, surrounded by a co-flow that may tions within the inner reaction zone may occur intermittently.
present different thermophysical properties (composition, temper- These results highlighted the variety of flame structures in the
ature, etc.). The flame is stabilized downstream of the injection and Jet Spray Flame configuration, associated to the burner geome-
presents several kinds of flame structures. Past pioneering works try and the fuel droplet generation. To enhance the understand-
ing of the droplets–turbulence–flame interactions and to improve
modeling capabilities of turbulent spray reacting flows, novel ex-

Corresponding author. perimental facilities were designed by several research groups to
E-mail address: renou@coria.fr (B. Renou).

https://doi.org/10.1016/j.combustflame.2018.03.032
0010-2180/© 2018 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
A. Verdier et al. / Combustion and Flame 193 (2018) 440–452 441

perform extensive measurements on the dispersed and carrier


phases.
The Sydney Spray Burner, developed by Masri and co-workers
[8–10], consists of a piloted spray burner where fuel is injected in
the carrier air co-flow by a nebulizer located upstream inside the
spray nozzle. The two-phase flame is stabilized by an annular pi-
lot flame. Authors provided an extensive database on the phenom-
ena of interest in dilute sprays: droplet dispersion, droplet evap-
oration, turbulence–droplets interactions and droplet–flame inter-
actions. The Delft Spray burner was designed by Roekaerts et al.
[11,12] to study spray combustion in Moderate or Intense Low-
oxygen Dilution (MILD) conditions. It consists of a pressure-swirl
atomizer that produces a spray of ethanol droplets issuing in a co-
flow of either air or hot combustion products. Dispersed and car- Fig. 1. Detail of the injection system. Dimensions in (mm).
rier phase properties were measured by Phase Doppler Anemome-
try (PDA), and a Coherent Anti-Stokes Raman Spectroscopy (CARS)
system was used to evaluate the gas-phase temperature statis- With the improvement of the high repetition rates, high-speed
tics. The Cambridge Spray Burner was updated from the gaseous diagnostics can resolve in time the scalars in turbulent flows.
bluff-body swirl burner by changing the injection system [13,14]. These efforts have yielded a greater understanding of the turbu-
N-heptane is injected through a pressure swirl hollow cone injec- lent flame dynamics by tracking the evolution in time of transient
tor located within the bluff-body surrounded by a swirled air co- phenomena such as local extinction, auto-ignition and turbulence–
flow. The burner operated close to blow-off limits and specific op- chemistry interactions [21–24]. However, an insight on the mech-
tical diagnostics (high-speed OH-PLIF and joint PLIF measurements anisms that control the flame dynamics and the transient phe-
of CH2 O and OH) were applied successfully both to investigate nomena such as extinction or re-ignition in two-phase flows is
the local flame structure, the reaction zones, and local extinction still in progress due to the measurement complexities in spray
holes along the flame sheet. More recently, the Rouen Spray Burner combustion. Despite the recent works on the application of high
was proposed as a new configuration to evaluate the droplet– speed diagnostics in spray combustion [13,25], obtaining accurate
flame interactions by using original optical diagnostics [15]. The experimental data on transient phenomena such as local extinction
experimental set-up is composed of an annular non-swirled air co- mechanisms in real and representative two-phase flow configura-
flow that surrounds a central hollow-cone spray injector, leading tions are still challenging.
to a flame stabilized downstream. Both PDA, OH-PLIF and Global In the current study, the CRSB database is completed with tem-
Rainbow Technique (GRT) were applied to quantify the evolutions poral and spatial measurements of the flame front and its associ-
of fuel droplet properties (size, velocity and temperature) across ated velocity field. Additional PDA results are also considered and
the flame front, leading to a complete database for validating LES presented. The objective of this work is to extend understanding of
codes. turbulence–flame interactions in the spray jet flame and droplet–
The previous databases appear to be fundamental tools to val- flame interactions by combining two high-speed optical diagnos-
idate or to improve LES codes, which are sensitive to the selected tics. This paper is organized as follows. Section 2 describes the
chemical mechanisms (global or detailed chemistry for instance) burner and the optical diagnostics. The global shape of the flame
and combustion models, and which are able to predict several and the associated aerodynamics in terms of dispersed and car-
types of local flame structure encountered in spray flame. Ma and rier phase are presented in the first part of Section 3. Moreover,
Roekaerts [16] performed a 3D LES with Flamelet Generated Man- the local extinction mechanisms which occur in the reaction zone
ifolds (FGM) of two configurations in the Delft Spray dataset. They will be discussed. A summary of the flame dynamics with transient
demonstrated that the flame structures are directly impacted by phenomena mechanisms is provided in Section 4.
any changes in the magnitude of different characteristic time scales
of the flow: fuel droplet evaporation, convection and reaction. The 2. Experimental setup
swirling ethanol spray flame near blow-off conditions (Cambridge
burner) was investigated by Guisti et al. [17,18] using the LES/CMC This section is dedicated to the description of the burner, the
approach with a simple and detailed chemical mechanisms. They operating conditions and the optical diagnostics applied during the
demonstrated that a detailed kinetic scheme was more appropriate experiments.
and able to capture the flame structure and its dynamics, such as
local extinctions and the lift-off height. The recent comparison be- 2.1. Burner and operating conditions
tween LES and the experimental results of the CORIA Rouen Spray
Burner (CRSB) database [15] was done by Shum-Kivan et al. [19]. Figure 1 illustrates the CORIA Rouen Spray Burner (CRSB),
The authors demonstrated the ability of LES (AVBP) to accurately which is based on the geometry of the gaseous KIAI burner [26].
predict the spray jet flame structure. However, although the global The atmospheric unconfined burner is composed of an external an-
shape is well predicted by the simulation, the position of the lead- nular non-swirling air co-flow and a pressurized liquid fuel injec-
ing edge is underestimated compared to the measurements de- tor (Danfoss, 1.35 kg h−1 , hollow cone with a angle of 80o ). The
duced from OH-PLIF. This demonstrates that the simulation of such air duct is equipped with 18 radial vanes (non-swirling) that break
geometries is a complex task, since it requires at least an accurate any large flow structures remaining in the plenum. The air-flow
forecast of the fuel vapor and thermal budget between the droplet is then guided by a convergent into the chamber surrounding the
and its gaseous surrounding through the evaporation model. A de- fuel injector and forming a turbulent air co-flow. An electronic
tailed description of the combustion reactions able to predict the Coriolis flow controller (Bronkhorst, CORI-FLOW, [0–2] g s−1 ) reg-
different modes of combustion is also necessary [20]. ulated the n-heptane mass flow rate and a thermal flow controller
Due to the complexity of the multi-physical phenomena, it is (Bronkhorst, EL-FLOW, [0–15] g s−1 ) was used to control the air
essential to develop and validate numerical models with accurate mass flow rate. The inlet conditions were 6 g s−1 (T = 298 ± 2 K)
experimental data for numerous practical combustion applications. and 0.28 g s−1 (T = 298 ± 2 K) for the air and fuel, respectively.
442 A. Verdier et al. / Combustion and Flame 193 (2018) 440–452

Fig. 2. Measurement grid for Phase Doppler Anemometry and fields of view for the
different optical diagnostics applied in the spray jet flame. To optimize the spatial
resolution, the field of view of the HS-OH-PLIF is smaller than the OH-PLIF.
Fig. 3. Optical arrangement for simultaneous HS-PIV and HS-OH-PLIF
measurements.
Cylindrical coordinates are considered: Uz and Ur represent the
axial and radial velocity respectively.
low stokes number. Both phases were measured in reactive condi-
2.2. Optical diagnostics tions. The size and velocity of the dispersed phase were also mea-
sured (in a different acquisition without seeding oil droplets) in
The local properties of the flow were measured by phase order to characterize the fuel droplet properties such as size dis-
Doppler anemometry (PDA) obtaining size-classified velocity data tribution and classified-size velocities. The PDA configuration and
for fuel droplets and velocity data for the air by seeding the car- the gain and voltage of the PMs were adjusted to detect all the
rier phase. The settings were adjusted in order to measure the dis- sizes present in the spray (2 ≤ D10 ≤ 80 μm) (configuration: PDA(B)).
persed and carrier phases in reactive conditions. The global flame Mean velocity and diameter values, presented in the next figures,
structure was investigated by OH-PLIF imaging at a low repetition were calculated with a minimum of 500 droplets. The velocity of
rate (10 Hz). High-Speed OH-PLIF (HS-OH-PLIF) and High-Speed each droplet was weighted by its transition time in the measure-
PIV (HS-PIV) were applied simultaneously. The temporal and spa- ment volume to avoid the well-known velocity bias.
tial flame front positions associated with instantaneous velocity
were obtained to study the transient phenomena such as extinc-
2.2.2. High resolution OH-PLIF (10 Hz)
tions. Figure 2 shows the different meshes and fields of view for
The global flame structure was investigated by OH-PLIF imag-
the different optical diagnostics.
ing at low repetition rate (10 Hz). A Nd:YAG.laser operating at
532 nm was used to pump a tunable dye laser (Quantel TDL90).
2.2.1. Phase doppler anemometry (PDA)
The excitation wavelength was tuned to the Q1 (5) transition of
A two-colour PDA measurement device from DANTEC was used  
the A2  + (v = 1 ) ← X 2 (v = 0 ) band of OH at λ = 282.665 nm.
to perform point measurements of air velocity and fuel droplet
The resultant output energy was 30 mJ per shot in the probe
velocity and size with green (514.5 nm) and blue (488 nm) laser
volume. The collection system consisted of an ICCD camera (PI-
beams. The emission probe used had a focal length of 350 mm and
MAX 4, Roper Scientific, 1024 × 1024 pixels2 ) equipped with UV
a beam spacing of 50 mm while the lens in the receiving probe
lens (f/2.8). The field of view was 112 × 112 mm2 leading to a mag-
had a focal length of 310 mm. The receiving probe was placed in
nification ratio of 4.52 pix.mm−1 . A broadband collection strategy
front scattering mode at 50o , close to the Brewster angle, enhanc-
from 308 to 330 nm with a band-pass filter (SCHOTT UG11) was
ing the detection of first order refracted light. A mask placed in
adopted and a high-pass filter (SCHOTT WG295) was used to re-
the receiver enabled a detection range of 139 μm. The beam cross-
duce Mie scattering from the fuel droplets.
ing generated a probe volume approximated by a cylinder 200 μm
long and with a diameter of 120 μm. Data acquisition was limited
to 30 s or 40,0 0 0 droplets so as to have converged statistics. Sev- 2.2.3. Simultaneous high-speed PIV and high-speed OH PLIF
eral parameters such as sphericity or droplet concentration control Temporal and spatial flame front positions and associated in-
the quality of the measurement. High-density regions lead to a de- stantaneous flow fields were obtained by High-Speed OH-PLIF at
crease in validation due to signal rejection when more than one 10 kHz combined with High-Speed PIV at 5 kHz. A schematic of
droplet at a time was present in the measurement volume. In this the high-speed imaging set-up is shown in Fig. 3. Concerning the
study, a decrease on the validation was observed below Z = 10 mm HS-OH-PLIF, a Nd:YAG.laser operating at 532 nm (104 W) was used
and it becomes unsatisfactory below Z = 7 mm in the dense region to pump a tunable dye laser (Sirah Credo). The excitation wave-

of the spray. Two different and independent acquisitions enabled length was tuned to the Q1 (5) transition of the A2  + (v = 1 ) ←

to focus the measurement strategy either on the air velocity or on X (v = 0 ) band of OH at λ = 282.671 nm. The resultant output
2

the fuel droplet size-conditioned velocity. Indeed, in order to cor- pulse energy was 380 μJ per shot in the probe volume. The beam
rectly detect the liquid and the gaseous phase separately, the gain was then expanded to 40 mm in height using a set of fused sil-
and voltage of the photomultipliers (PMs) were accordingly ad- ica lenses ( f1 = 10 0 0 mm, f2 = −20 mm, f3 = 500 ) mm. Due to the
justed. To perform the measurements of the carrier phase in pres- low energy delivered, the detection system consisted of a CMOS-
ence of fuel spray, the air was previously seeded with 2 μm olive camera Photron Fastcam SA5 mounted with an external image in-
oil droplets and the gain and voltage in the PMs were optimized tensifier (High Speed IRO, LaVision). The camera was operated at
(increased) in consequence (configuration: PDA(A)). PDA(A) config- a repetition rate of 10 kHz with an array of 896 × 848 pix2 and a
uration allows to detect the air seeding and small fuel droplets magnification ratio of 20.41 pix.mm−1 . The intensifier delay (26 ns)
D10 ≤ 2 μm, which follow perfectly the air velocity due to their and gate (500 ns) were set to optimize the signal to noise ratio.
A. Verdier et al. / Combustion and Flame 193 (2018) 440–452 443

Fig. 4. (a) Photograph of the spray jet flame. (b) Typical OH-PLIF recording and different flame zones (A, Inner Reaction Zone (IRZ), C, Outer Reaction Zone (ORZ) and S).

Two high-pass optical filters (SCHOTT WG295) were used to re- flow rate, spray properties (droplet velocity and size) and global
duce background noise due to elastic scattering by the droplets. equivalence ratio, as it was reported in many papers. The axi-
The camera on-board memory can hold over 7500 frames, corre- symmetric flame with a classical double structure is illustrated by
sponding to an acquisition time of 750 ms. The OH-PLIF signal was the OH-PLIF (Fig. 4(b)), where five different zones (A, IRZ, C, ORZ
collected within the 308–330 nm range using a band-pass filter and S) are reported. The overall shape of the flame exhibits two
(SCHOTT UG11). Moreover, the laser sheet profile was taken into branches corresponding to an inner and an outer reaction zone, IRZ
account and corrected by filling a quartz vessel placed on the test and ORZ respectively, connected together at the flame leading edge
facility with a homogeneous air–acetone mixture. Image process- (S). These two reaction zones are separated by hot gases (zone C)
ing tools including non-linear filtering [27] and active contour de- exhibiting a weak OH-PLIF signal. This flame structure results from
tection implemented in a level-set method [28] were developed to the spray heterogeneity in size [5,12] imposed by the pressure fuel
extract flame front contours. The method is based on the original injector technology. The large droplets spread into the outer part
approach formulated by Perona and Malik [29] and has several ad- of the air co-flow corresponding to the ambient air and the small
vantages: noise is smoothed locally, whereas little or no smooth- droplets are mainly located near the centerline in the mixing zone
ing occurs between image objects which enhance the local edges (A).
[27,30]. Moreover, non-linear diffusion filter allows preserving the
flame front edges without artificial shifting in position [31]. 3.2. Carrier and dispersed phases properties
To obtain instantaneous velocity fields, high speed PIV was
applied at 5 kHz. A double cavity (Nd:YLF Laser Darwin Dual Figure 5 presents a map of the two mean velocity components
Quantronix) operating at 527 nm delivers two laser beams sep- (axial and radial) of the carrier phase obtained by PDA in con-
arated by tPIV = 10 μs. Zirconium oxide (ZrO2 ) particles were figuration PDA(A). Two important features are illustrated in the
used to seed the air co-flow (configuration: HS-PIV-(A)). The Mie mean velocity fields. Consistent with the earlier works [2,7,11,25],
scattering signal from droplets and ZrO2 was collected with a the mean axial co-flow velocity provides two peaks at the exit of
Phantom V2512 camera (1280 × 800 pix2 ) equipped with a Nikon the injector. For high axial stations Z > 25 mm, the bluff-body ef-
f/1.4 50 mm objective leading to a magnification ratio equal to fect becomes negligible and the co-flow air peaks merge into a flat
18.98 pix.mm−1 . In this study, a close attention was paid to the axial velocity profile. Note that the air co-flow jet presents radial
air entrainment by the air co-flow. To capture the slow air entrain- velocities approximately 5 times smaller than the axial velocities
ment velocity, the HS-PIV was adjusted (configuration: HS-PIV-(B), at Z = 10 mm, which rapidly decrease down to zero at Z = 25 mm.
tPIV = 400 μs). The ambient air was seeded with a glycol mist in Magnitude of velocity fluctuations is represented in Fig. 6 in
order to obtain the air entrainment, which is crucial to explain the terms of the turbulent kinetic energy k. Isotropy along the two di-
flame structure. The same post-treatment (Dynamics Studio soft- rections perpendicular to the Z-axis is assumed and turbulent ki-
ware) was applied on images (HS-PIV-(A) and HS-PIV-(B)) with a netic energy (k) is calculated following equation:
cross-correlation algorithm with a window size of 32 × 32 pix2 and  
1
a 50% overlap. A coherent filter was applied to correct isolated k= (uz )2 + 2 × (ur )2 (1)
false vectors. 2
The annular air co-flow discharging into the ambient air cre-
3. Results and discussion ates a shear layer at the interface between the high air co-flow
velocity and the ambient air inducing turbulence with large scale
This section is divided into four subsections that correspond to structures. This phenomenon is highlighted in the movie given in
the results of the global analysis of aerodynamics, flame shape, po- Supplementary Material SM1. The quiescent ambient air is accel-
sition of local extinctions and the study of interactions between erated radially towards the centerline whereas a portion of the air
flame, turbulence and droplets. jet loses some momentum. Figure 6(b) represents different turbu-
lent kinetic energy profiles for three axial stations. The maximum
3.1. Global topology of the spray jet flame value of k (≈ 40 m2 s−2 ) is found at low axial stations, in the region
where the air enters the chamber. The turbulent kinetic energy
Figure 4(a) shows a photograph of the spray jet flame, where does not follow the same pattern as the axial velocity, with peaks
the global flame appearance can be observed. The global flame diffusing radially with increasing axial stations. Given that velocity
shape and the lift-off height of the flame are governed by air fluctuations enhance mixing and that all droplets enter the region
444 A. Verdier et al. / Combustion and Flame 193 (2018) 440–452

Fig. 5. Mean components of air velocity flow for reacting conditions with vector field obtained with configuration PDA(A). (a) Mean axial velocity. (b) Mean radial velocity.

Fig. 6. (a) Turbulent kinetic energy (k) of the air in reacting conditions, calculated
from Eq. 1. (b) Turbulent kinetic energy profiles for different axial stations.

of maximum k directly after atomization, the small droplet trajec-


tories will be strongly influenced by the airflow velocities. This in-
teraction segregates the droplets spatially, transporting the small
droplets (with low Stokes number) to the center while the large
droplets are directed to the spray borders. The fuel spray shows
a complex heterogeneous distribution, with droplet sizes ranging
from 2 to 60 μm. Profiles of the mean diameter (D10 ) and Sauter
mean diameter (SMD, D32 ), at four axial stations, Z = 16, 25, 35 and
45 mm, are represented in Fig. 7(a) and (b). The mean diameter
profiles grow (from 5 to 50 μm) radially from the center towards Fig. 7. (a) Fuel droplet Sauter Mean Diameter (SMD) D32 in reacting condition. (b)
Fuel droplet mean diameter D10 in reacting condition. Error bars indicate the RMS
the borders and vary significantly with Z due to the evaporation
of the droplet size distribution. (c) PDF of droplet size fitted with Rosin-Rammler
process. Droplets are expelled from the atomizer under a cone distribution.
with an 80o opening, following a direction that will force them
to suddenly interact with the air co-flow. The central region con-
tains a very low quantity of droplets due to the centrifugal move- To illustrate the behavior of the droplets, the mean and RMS
ment imposed by the swirling liquid injection inside the atomizer. of the axial and radial velocity components of different droplet
The smallest droplets resulting from the liquid jet breakup are re- size-classes at four axial stations, Z = 16, 25, 35 and 45 mm,
sponsive to the air velocity field and follow the flow due to their are shown in Figs. 8 and 9. The class width was chosen to be
associated small Stokes numbers. Furthermore, in this region, the 10 μm. The behavior of the 60–70 size group is very close to
very small droplets represent only a small quantity of the liquid that of the 50–60 group but, since they contain less droplets,
mass leading to a lean heterogeneous mixture that is close to the they are not presented. Values for the last two groups (30–40,
lean flammability limit [19]. However, larger droplets with higher 50–60) are not shown in the central region because less than
Stokes numbers are less affected by the aerodynamic conditions 500 droplets in 30 s were detected per point. The carrier phase
and follow ballistic trajectories. It is also worth noting that the velocity profiles are also superimposed to evaluate slip velocity
number distributions of droplet size are well represented by Rosin- statistics. Figure 8 shows that droplets are ejected from the nozzle
Rammler distributions everywhere, as illustrated in Fig. 7(c). with strong radial velocities. At the first axial station, the three
A. Verdier et al. / Combustion and Flame 193 (2018) 440–452 445

Fig. 8. Mean components of fuel droplet velocity in reacting conditions separated in four size-classes obtained with PDA (B) configuration. o represent the [0–10] μm group,
the [20–30] μm group, the [30–40] μm and represent the [50–60] μm. Black lines represent the Air velocities profiles.

Fig. 9. RMS of fuel droplet velocity in reacting conditions separated in four size-classes. o represent the [0–10] μm group, the [20–30] μm group, the [30–40] μm and
represent the [50–60] μm. Black lines represent the Air RMS velocities profiles.

groups have appreciable radial velocities (≈ 10 m s−1 ) but farther action Zone topology. The IRZ exhibits a highly wrinkled struc-
downstream small droplets adapt their velocity to the airflow ture placed on the shear layer formed between the issuing air jet
velocities. The large droplets follow more ballistic trajectories and and the surrounding air. It was shown numerically that IRZ ex-
continue to have radial centrifugal velocities until Z = 35 mm. hibits a double (premixed and non-premixed) structure [19]. In-
When overlapped to the air stream, axial mean velocities of small deed, the IRZ consists first in a premixed flame fed by the cen-
droplets (groups 0–10 μm) are similar to that of the air whereas tral part of the spray (Zone A) where the mixing between small
large droplets continue moving at smaller velocities until farther droplets (D10 < 20 μm) and the air co-flow generates a very lean
downstream. When fuel droplets exit the high air velocity region, region close to the flammability limit of n-heptane. Large amount
they find a quiescent area where they can decelerate. Concerning of fuel vapor is found between the Inner and Outer reaction zones,
the velocity fluctuations, small droplets are more affected by where medium to large droplets (D10  20–60 μm), which account
the shear layer and the coherent structures generated by the air for the major part of the injected fuel mass, evaporate fast result-
co-flow turbulence (Fig. 9). Near the radial station R = −10 mm, ing in a fuel rich mixture. Consequently, the remaining air from
in the region where the turbulent kinetic energy is maximum and the lean premixed flame then reacts with the fuel vapor that dif-
for axial stations where large turbulent scales are fully developed, fuses from the fuel-rich, hot region (zone C) in a diffusion flame
the RMS for radial velocities is maximum. very close to the lean premixed flame. The IRZ is then located in a
very lean region of the flow (for its inner flame) and subjected to
3.3. Flame structures strong velocity fluctuations. Consequently, the IRZ is strongly dis-
turbed (momentum and mass transfer) and wrinkled by local ve-
3.3.1. Inner reaction zone (IRZ) locity gradients, and frequent events of transient phenomena such
An instantaneous snapshot of OH-PLIF at 10 Hz in a vertical as local extinctions are observed. These local extinctions are illus-
cross section is reported in Fig. 4(b) and illustrates the Inner Re- trated by holes in High-Speed OH-PLIF video given as supplemen-
446 A. Verdier et al. / Combustion and Flame 193 (2018) 440–452


Fig. 10. (a) Spatially averaged ORZ position (black lines obtained with OH-PLIF measurements) and air entrainment streamlines with the norm of velocity (|U | = U2 + V 2)
obtained with HS-PIV-(B) configuration). (b) Schematic representation of ORZ : Generic structure of a laminar diffusion flame.

tary material 2 and will be investigated in more details in the fol-


lowing section.

3.3.2. Outer reaction zone (ORZ)


Fuel droplets both in non-reactive zones (spray exit, mixing
zone) and in reactive zones (zones C and ORZ) can be easily iden-
tified by Mie scattering in Fig. 4(b). The ballistic trajectories of
large inertial droplets lead to the presence droplets in zone C,
where the high temperature increases the evaporation rate gener-
ating a gaseous fuel reservoir. Despite there are few big droplets
(D10 > 30 μm) coming to the outer region of the spray, the droplet
evaporation rate implies the production of fuel vapor until reach-
ing an equivalence ratio higher than φ = 1.8 [19]. In addition,
fuel thermal decomposition and incomplete hydrocarbon combus-
tion will produce CO [16], CO2 and H2 O, and intermediate species
Fig. 11. Positions of the flame base.
such as soot precursors (PAH), as evidenced by the yellow color
of the flame. Consequently, the fuel mass fraction YF is obviously
far less than unity in zone C, and the mixture fraction falls be-
tween ξ st < ξ < 1. The air entrainment is shown in Fig. 10(a), where
streamlines and the velocity magnitude are illustrated. The stream-
lines are uniform far away from the air jet, and diverge close to the
ORZ with a stagnation point located in average at R = 37 mm and Z
= 52 mm. At this stagnation point, flow trajectories are separated:
one move towards the leading edge and the second one towards
the top of the flame. It is worth noting that the quiescent sur-
rounding air is accelerated towards the air co-flow exit and reaches
1 m s−1 just below the flame leading edge. The entrained air re-
acts with the gaseous fuel reservoir in a diffusion-like outer flame
front (ORZ), where a stoichiometric line is present and located be-
tween the rich (partially burned gases) and lean (fresh gases) sides.
Compared to the IRZ, the ORZ is less wrinkled, more stable, thicker
and characterized by a smoother OH gradient. Figure 10(b) sum-
Fig. 12. Mean OH-PLIF with extinction locations. Red inverted triangles represent
marizes the ORZ structure.
the local flame extinction positions in the IRZ, whereas blue triangles concern the
local extinction positions in the flame leading edge. Green ellipse mark a region,
3.3.3. Stabilization of the flame base (S) where large inertial droplet cross the flame leading edge and interact with the ORZ.
The stabilization mechanism of a lifted spray jet flame is con- (For interpretation of the references to color in this figure legend, the reader is
ditioned by the droplet convection-vaporization, the flow aerody- referred to the web version of this article).
namics and the chemical reaction budgets. For this specific op-
erating condition, the flame is lifted and stabilized around R =
15.5 ± 3 mm and Z = 24 ± ± 3 mm. Each square in Fig. 11 This preliminary description of the different zones of the Jet
corresponds to the leading edge location extracted from a set of Spray Flame illustrate the strong interactions between large inertial
500 OH-PLIF images. This position presents a favorable condition to droplets and the leading edge revealing also some local extinctions
sustain a stable reaction: the turbulence is low (Fig. 6) and there is in the IRZ and on the flame leading edge. The time-scales of tur-
enough fuel vapor available fed by the evaporation of large inertial bulence and chemistry are small and require accurate experimen-
droplets. tal data through high-speed measurements to analyze the transient
A. Verdier et al. / Combustion and Flame 193 (2018) 440–452 447

Fig. 13. Extinctions in the IRZ : flame–turbulence interactions. HS-OH-PLIF images overlaid with the fluctuation velocity field and the strain rate (SM3) - time separation =
0.2 ms.

phenomena that occur in a spray jet flame. This is presented in the three different zones of local extinctions associated with three dif-
following sections. ferent flame structures. First, local flame extinctions are observed
in the IRZ (red inverted triangles) in axial regions from R = 10
3.4. Position of local extinctions and R = 15 mm for different heights above 27 mm. These local
extinctions occurs the most frequently in a region where the flame
In this work, OH-PLIF signal is used as a marker of the reaction front is submitted to intense local strain. Occasionally, some local
zones since disruptions in continuous OH fronts correlate well extinctions occur close to the flame leading edge (blue triangles).
with local flame front extinctions [32]. Here, the different zones This zone is characterized by large droplets (D10 > 30 μm) with a
of local flame extinction are marked in Fig. 12 where they are mean velocity equal to 10 m s−1 . At this location, the local turbu-
superimposed with the mean OH-PLIF image. These markers lence intensity is moderate, and the extinction is therefore mainly
represent the position (R, Z) of local flame extinction appearance controlled by droplets–flame interactions. The third zone concerns
extracted from time series of HS-OH-PLIF images. One can notice the ORZ, where the large initial droplets expelled from the pres-
448 A. Verdier et al. / Combustion and Flame 193 (2018) 440–452

Fig. 14. Schematic representation of the extinction mechanisms in the inner reaction zone.

Fig. 15. Profile at R = 16 mm and Z = 30 mm. (a) Extinction occurrence temporal analysis. (b) Extinction occurrence frequency analysis with three runs.

sure injector cross the flame leading edge and are still present ing velocity fields obtained by HS-PIV. The OH signal is associated
when they interact with the ORZ. The perturbations between large to both reaction zones (premixed and non-premixed) as it was
droplets and ORZ (which are not extinctions) are observed in the numerically demonstrated in [19] producing OH radicals in IRZ.
green delimited region in Fig. 12. However, these reaction zones are very close together and it
In summary, three regions of transient phenomena are observed is not experimentally possible to discriminate OH signals from
in the spray jet flame and defined here for subsequent analysis their origin. A clear disruption in continuous OH front in the IRZ
with more details: means that OH signal is neither observed from premixed nor
(i) Extinctions in the IRZ far away from the flame leading edge: non-premixed reaction zones. Note that the velocity vectors are
flame–turbulence interactions. overlaid with the instantaneous strain rate of flow and computed
 1
(ii) Extinctions in the IRZ close to the flame leading edge: as σ = ( Ei2j ) 2 , where Eij is estimated by Ei j = 12 (∂ j vi + ∂i v j ), the
droplets–flame interactions.
2D strain rate tensor. The maximum value of σ is around 50 0 0 s−1
(iii) Perturbations in the ORZ: droplets–flame interactions.
and appears intermittently in the flow region interacting with the
IRZ. These large scales of velocity fluctuations are created by the
3.5. Flame–turbulence interactions
velocity gradients in the shear layer between the co-flow and the
ambient air (Supplementary Materials, SM3). Figure 13 highlights
Figure 13 presents a sequence of flame extinctions in the IRZ
that the flame front thinning and wrinkling are caused by a high
illustrated by OH-PLIF images superimposed with the correspond-
A. Verdier et al. / Combustion and Flame 193 (2018) 440–452 449

Fig. 16. HS-OH-PLIF images - time separation = 0.1 ms. Extinction in the leading edge due to the droplet–flame interactions.

Fig. 17. HS-OH-PLIF images - time separation = 0.1 ms. Perturbations in the outer reaction zone (ORZ) due to the droplet–flame interactions.

strain rate (σ > 50 0 0 s−1 ) in IRZ region. From t = 0.0 ms to t a local extinction in the IRZ cannot cause a global extinction of the
= 0.4 ms the OH signal in the flame front becomes thinner and flame.
thinner. The extinction event is observed on the HS-OH-PLIF im- Figure 14 presents a schematic representation of the
ages at t = 0.4 ms, which correspond to a position R = 12 mm and turbulence–flame interactions in the IRZ. Close to the injector
Z = 35 mm. The local extinction seems to appear when the flame when the co-flow is discharging in the quiescent air, large coher-
front is too fine to support a high stretch rate. Then, with the local ent and intense turbulence structures are created by the mean
extinction in the IRZ, a gap is produced and causes a hole separat- velocity gradient. They are large enough to go far out into the
ing the inner flame front in two reaction zones. The hole and the co-flow and into zone (C) and then to induce a strong large scale
local IRZ are convected downstream due to the mean air velocity mixing between fuel droplets, vapor and air. Then, this local
field. The hole propagation rate is not constant and depends on the heterogeneous and turbulent mixture interacts with the reaction
local turbulent flow and composition. Note that in this configura- zone and disturbs the flame front by wrinkling processes. In
tion, and in the OH-PLIF cross section, 11 local extinctions in the the hot region close to the IRZ the kinematic viscosity changes
IRZ are observed for 10 0 0 instantaneous OH-PLIF images. More- approximately by a factor of 20 from fresh to burned gases. Even
over, with these operating conditions (air: 6 g s−1 ; fuel: 0.28 g s−1 ) if the flow is laminarized by the higher viscosity close to the
450 A. Verdier et al. / Combustion and Flame 193 (2018) 440–452

IRZ, the flame is disturbed by the flow aerodynamics. When local


strain rate associated with local equivalence ratio exceeds a critical
value, local flame extinction occurs and the extinguished part
is convected downstream by the flow field. Indeed, recent study
[33] demonstrates that during flame–vortex interactions, the flame
can locally extinguish when the flow exceed a critical extinction
strain rate. The main difference with local extinction with gaseous
flames [34,35] is on the presence of small fuel droplets near the
reaction zone which can enhance fuel heterogeneities or quickly
evaporate when crossing the flame front. Franzelli et al. [36] has
shown that local extinctions occurring in a spray jet flame may
also be induced by local fuel depletion.
To validate the predominant aerodynamic effect on local extinc-
tion mechanisms in the IRZ, caused by local velocity gradients and
lean mixtures, an extinction frequency analysis is applied to dif- Fig. 18. Schematic representation of the droplet–flame interactions.
ferent HS-OH-PLIF measurements. An analysis region with a probe
located in the mixing region in the vicinity of the reaction zone
(R = 14 mm and Z = 30 mm) is used to determine the temporal ap- local flame extinction. A detailed analysis of the OH-PLIF images
pearance of extinctions. In Fig. 15(a), Df represents the distance be- was performed to identify the number of extinctions during a set
tween the probe and the flame front position, extracted from OH- of measurement. In the plane of measurement, 91 local extinctions
PLIF images. The specific value D f = 0 corresponds to the position events were detected for 10 0 0 instantaneous OH-PLIF images or for
of the instantaneous flame front when the IRZ is located at the 100 ms.
same position as the probe, whereas the negative and positive val- Previous numerical studies [37,38] suggest that the droplet–
ues of Df imply a flame-probe distance located in the fresh and flame interaction mechanisms, in a partially premixed flame,
burnt gases, respectively. The values of D f > 10 mm correspond to should be related to a thermal disturbance and to a modification of
the distance between the probe and the ORZ. This occurs when the local mixture. Indeed, Nakamura and Akamatsu [37] show that
there is a local extinction in the IRZ. The temporal evolution of the energy exchange between droplets and gaseous phase may dis-
Df has a so-called telegraphic shape, alternating D f > 10 mm and turb the reaction rate. The temperature of droplets is much lower
D f  0 mm values. The power spectral density function (PSD) of than of the gaseous phase, acting as a sink term for the chemical
the temporal evolutions of Df are reported in Fig. 15(b). A strong reactions. The intermittent passage of large droplets in the leading
peak around 80 Hz is clearly visible, repeatable, indicating pe- edge induces a locally high strain rate due to the droplet motion,
riodic flame extinction phenomenon. These measurements have which can reduce the reaction rate [39–41]. When crossing the
been performed three times separated by several weeks. It can be flame, fuel droplet temperature increases up to a welt-bulb tem-
observed that for the red and blue lines, the PSD show a strong perature [15], which remains lower than the fuel boiling tempera-
peak around 80 Hz whereas for the third acquisition (green) a peak ture. In addition, the high burned gas temperature leads to a strong
at around 80 Hz is still observed but with a more broadened shape. fuel evaporation rate and a large values of gaseous fuel mass frac-
The frequency analysis of the large-scales of turbulence seems tion YF gas around the droplets. The local extinction in the leading
relevant for the understanding of the extinction mechanisms and it edge can be explained by the droplet temperature in this region
is now discussed. From high-speed PIV measurements, a temporal (Tdroplet Tflame ) and the mixture around the droplet, which is very
velocity signal is extracted at R = 13 mm and Z = 30 mm located rich and outside the flammability limits. It is worth noting that
in the shear layer (region A). The integral of the auto-correlation the local flame extinction due to droplets crossing a flame front is
function of these velocity fluctuations is then determined and is not explicitly taken into account in most models (closure models
equal to 12.2 ms. The latter represents the temporal integral length in RANS or subgrid models in LES).
scale of the turbulence at this location, where local flame extinc-
tions occur. It represents a measurement of the longest time that 3.6.2. Perturbations in the ORZ
the velocity fluctuation remains autocorrelated being, hence, rep- Figure 17 presents HS-OH-PLIF images separated by 0.1 ms,
resentative of coherent large structures in the flow. In this case it where two big droplets (red circles) cross the zone C with a slip
is equivalent to a frequency of 82 Hz. This value is similar to the velocity of ≈ 7 m s−1 . Further downstream, droplets that have sur-
flame extinction occurrence underlying that flame extinction in IRZ vived the leading edge move towards the ORZ, where they disturb
is mainly controlled by flame/turbulence interactions. Indeed, in the flame front. Indeed, after the passage of the drops, a local per-
the mixing zone of the flame (zone A), the lean mixture is very turbation located in the wake can be easily identified. The movie
close to the flammability limit [19] and the local flame strain rate of this droplet–flame interaction is available in Supplementary Ma-
induced by the most energetic length scales of the shear layer will terials (SM2).
exceed locally and periodically the extinction strain rate of the IRZ. Figure 18 presents a schematic representation of the droplet–
flame interaction. As described in the previous section, large
3.6. Droplet–flame interactions droplets cross the leading edge and continue to evaporate in
burned gases in zone C. The strong thermal exchange leads to a
3.6.1. Extinctions in the leading edge fast evaporation regime resulting in an increase of the gaseous fuel
As already seen in the previous section, the small droplets mass fraction and, hence, a droplet diameter reduction. However,
mainly stay in the cold central zone, where they evaporate slowly the largest droplets can reach the ORZ and induce local pertur-
and never cross the flame front. On the contrary, large inertial bations in the reaction zone. In the burned gases (zone C), lot of
droplets (D10 > 30 μm) move to the external part of the spray cross- species are present and imply gaseous fuel mass fraction YF < 1.
ing the flame base and reaching the hot region. Figure 16 presents However, when considering fuel droplets within the ORZ, the pure
a local extinction event in the leading edge caused by a large fuel vapor generated by the evaporation leads to an increase of the
droplets. At t = 0.2 ms, a large droplet (yellow circle) crosses the fuel mass fraction until to reach YF close to one near the liquid in-
leading edge. 0.1 ms after, a hole is produced corresponding to a terface. The YF fluxes towards the reaction zone are then strongly
A. Verdier et al. / Combustion and Flame 193 (2018) 440–452 451

modified in the vicinity of the droplets and should locally mod- Acknowledgments
ify the location of the stoichiometric line (ξ St ) in the ORZ. The
droplet evaporation decreases also the surrounding gas tempera- The authors gratefully acknowledge the financial support from
ture within the ORZ due to the heat latent absorption until reach- the Agence Nationale de la Recherche with the grant TIMBER
ing the wet-bulb temperature of the droplet. This latter was mea- ANR-14-CE23-0 0 09. Dr. Malbois and Salaun are also warmly ac-
sured to around 330 K in zone C by Verdier et al. [15] and may knowledged for their technical support on data processing tools.
locally quench the flame. The droplet–flame mechanism in the ORZ
is then related to a cooling effect (strong gradient between the
Supplementary material
droplet and the flame temperature Tdroplet Tflame ) and to a lo-
cal increase of the gaseous fuel mass fraction around the droplet,
Supplementary material associated with this article can be
leading to the appearance of a wake behind the droplets traveling
found, in the online version, at 10.1016/j.combustflame.2018.03.
across the ORZ. In most cases, the droplets disturb the flame front
032.
but no extinctions can be observed, different from what is seen in
the premixed flame zone (IRZ). Furthermore, the perturbation time
is shorter than the local extinction duration in IRZ. In the ORZ, the References
turbulence intensity is very low and combustion occurs in diffu-
[1] P. Goix, C. Edwards, A. Cessou, C. Dunsky, D. Stepowski, Structure of a
sion regime, which prevents the amplification of the perturbation methanol/air coaxial reacting spray near the stabilization region, Combust.
caused by the droplets. Flame 98 (3) (1994) 205–219.
[2] A. Cessou, D. Stepowski, Planar laser induced fluorescence measurement of hy-
droxyl radical in the stabilization stage of a spray jet flame, Combust. Sci. Tech-
nol. 118 (4–6) (1996) 361–381.
[3] D. Stepowski, P. Goix, A. Cessou, Simple description of the combustion struc-
4. Conclusion tures in the stabilization stage of a spray jet flame, At. Sprays 9 (1) (1999)
1–27.
A detailed experimental study of turbulence–droplet–chemistry [4] J.A. Friedman, M. Renksizbulut, Investigating a methanol spray flame inter-
acting with an annular air jet using phase-doppler interferometry and planar
interaction events in an n-heptane spray jet flame has been pre- laser-induced fluorescence, Combust. Flame 117 (4) (1999) 661–684.
sented. The sequences of the simultaneous high-speed measure- [5] S.K. Marley, E.J. Welle, K.M. Lyons, W.L. Roberts, Effects of leading edge en-
ments allow to capture the transient phenomena, such as the trainment on the double flame structure in lifted ethanol spray flames, Exp.
Thermal Fluid Sci. 29 (1) (2004a) 23–31.
mechanisms governing the local extinctions events with sufficient [6] S.K. Marley, K.M. Lyons, K. Watson, Leading-edge reaction zones in lifted-jet
spatial and temporal resolution. The flame exhibits a double struc- gas and spray flames, Flow Turbul. Combust. 72 (1) (2004b) 29–47.
ture with inner and outer reaction zones, where fuel droplets are [7] G. Cleon, D. Honore, C. Lacour, A. Cessou, Experimental investigation of struc-
ture and stabilization of spray oxyfuel flames diluted by carbon dioxide, Proc.
still present and interact with the reaction zones. Each flame zone
Combust. Inst. 35 (3) (2015) 3565–3572.
exhibits different mechanisms of droplet–turbulence–flame inter- [8] S.H. Starner, J. Gounder, A.R. Masri, Effects of turbulence and carrier fluid on
actions. simple, turbulent spray jet flames, Combust. Flame 143 (4) (2005) 420–432.
[9] A.R. Masri, J.D. Gounder, Turbulent spray flames of acetone and ethanol ap-
proaching extinction, Combust. Sci. Technol. 182 (4–6) (2010) 702–715.
• Flame/Turbulence extinction mechanism. Periodic local flame ex- [10] J.D. Gounder, A. Kourmatzis, A.R. Masri, Turbulent piloted dilute spray flames:
tinctions are observed in the IRZ downstream the leading edge flow fields and droplet dynamics, Combust. Flame 159 (11) (2012) 3372–3397.
[11] H. Correia Rodrigues, M.J. Tummers, E.H. van Veen, D.J.E.M. Roekaerts, Spray
for axial stations above Z = 27 mm. These local extinctions oc- flame structure in conventional and hot-diluted combustion regime, Combust.
cur in a region where the lean partially vaporized premixed Flame 162 (3) (2015a) 759–773.
flame front is submitted to intense local strain rate induced by [12] H. Correia Rodrigues, M.J. Tummers, E.H. van Veen, D.J.E.M. Roekaerts, Effects
of coflow temperature and composition on ethanol spray flames in hot-diluted
the shear layer. This extinction mechanism is mainly associated coflow, Int. J. Heat Fluid Flow 51 (2015b) 309–323.
to flame–turbulence interactions in the gas phase since most of [13] D.E. Cavaliere, J. Kariuki, E. Mastorakos, A comparison of the blow-off be-
the small droplets within this region have been rapidly evapo- haviour of swirl-stabilized premixed, non-premixed and spray flames, Flow,
Turbul. Combust. 91 (2) (2013) 347–372.
rated at the entrance of the reaction zone. [14] R. Yuan, J. Kariuki, A. Dowlut, R. Balachandran, E. Mastorakos, Reaction zone
• Flame/Droplet extinction mechanism. The mechanism related to visualisation in swirling spray n-heptane flames, Proc. Combust. Inst. 35 (2)
the droplet–chemistry interactions occurs at the flame leading (2015) 1649–1656.
[15] A. Verdier, J. Marrero Santiago, A. Vandel, S. Saengkaew, G. Cabot, G. Grehan,
edge and occasionally produces a local flame extinction. In-
B. Renou, Experimental study of local flame structures and fuel droplet prop-
deed, this zone is characterized by low turbulence intensities erties of a spray jet flame, Proc. Combust. Inst. 36 (2) (2017) 2595–2602.
and large droplets crossing the flame front with high velocity. [16] L. Ma, D. Roekaerts, Numerical study of the multi-flame structure in spray
combustion, Proc. Combust. Inst. 36 (2) (2017) 2603–2613.
These droplets are characterized by a low temperature thus act-
[17] A. Giusti, M. Kotzagianni, E. Mastorakos, LES/CMC simulations of swirl-sta-
ing as a heat sink for the flame front, which finally extinguishes bilised ethanol spray flames approaching blow-off, Flow, Turbul. Combust. 97
mainly due to the cooling effect. (4) (2016) 1165–1184.
• Flame/Droplet perturbation mechanism. The last mechanism as- [18] A. Giusti, E. Mastorakos, Detailed chemistry LES/CMC simulation of a swirling
ethanol spray flame approaching blow-off, Proc. Combust. Inst. 36 (2) (2017)
sociated to the droplet–flame interaction mechanism occurs 2625–2632.
in the ORZ. The largest droplets reaching the ORZ evaporate [19] F. Shum-Kivan, J. Marrero Santiago, A. Verdier, E. Riber, B. Renou, G. Cabot,
quickly in the burnt gases region producing large amount of B. Cuenot, Experimental and numerical analysis of a turbulent spray flame
structure, Proc. Combust. Inst. 36 (2) (2017) 2567–2575.
fuel vapor surrounding the droplet. They interact with the dif- [20] L. Ma, B. Naud, D. Roekaerts, Transported pdf modeling of ethanol spray in
fusion flame front where a strong decrease of the OH is ob- hot-diluted coflow flame, Flow, Turbul. Combust. 96 (2016) 469–502.
served around the droplet during its trajectory. If no local ex- [21] I. Boxx, C.M. Arndt, C.D. Carter, W. Meier, High-speed laser diagnostics for the
study of flame dynamics in a lean premixed gas turbine model combustor, Exp.
tinction is observed, a strong modulation of OH signal indicate Fluids 52 (3) (2010) 555–567.
a local decrease of the flame heat release rate. [22] M. Stohr, R. Sadanandan, W. Meier, Experimental study of unsteady flame
structures of an oscillating swirl flame in a gas turbine model combustor, Proc.
Combust. Inst. 32 (2) (2009) 2925–2932.
All these experimental results provide new insight on the [23] C. Abram, B. Fond, A.L. Heyes, F. Beyrau, High-speed planar thermometry and
experimental evidences of droplets/flame/turbulence interaction velocimetry using thermographic phosphor particles, Appl. Phys. B 111 (2)
mechanisms in a spray jet flame, and an original database suit- (2013) 155–160.
[24] C.D. Slabaugh, A.C. Pratt, R.P. Lucht, Simultaneous 5 kHz OH-PLIF/PIV for the
able for numerical modeling to predict non-stationary events such study of turbulent combustion at engine conditions, Appl. Phys. B 118 (1)
as local flame extinction in spray flames. (2015) 109–130.
452 A. Verdier et al. / Combustion and Flame 193 (2018) 440–452

[25] M.S. Mansour, I. Alkhesho, S.H. Chung, Stabilization and structure of n-heptane [33] M.K. Geikie, K.A. Ahmed, Lagrangian mechanisms of flame extinction for lean
flame on CWJ-spray burner with kHZ SPIV and OH-PLIF, Exp. Thermal Fluid Sci. turbulent premixed flames, Fuel 194 (2017) 239–256.
73 (2016) 18–26. [34] P.H. Renard, D. Thevenin, J.C. Rolon, S. Candel, Dynamics of flame/vortex inter-
[26] M. Cordier, A. Vandel, G. Cabot, B. Renou, A.M. Boukhalfa, Laser-induced spark actions, Progr. Energy Combust. Sci. 26 (3) (20 0 0) 225–282.
ignition of premixed confined swirled flames, Combust. Sci. Technol. 185 (3) [35] A.M. Steinberg, J.F. Driscoll, Straining and wrinkling processes during turbu-
(2013) 379–407. lence–flame interaction measured using temporally-resolved diagnostics, Com-
[27] H. Malm, G. Sparr, J. Hult, C.F. Kaminski, Nonlinear diffusion filtering of images bust. Flame 156 (12) (2009) 2285–2306.
obtained by planar laser-induced fluorescence spectroscopy, J. Opt. Soc. Am. A [36] B. Franzelli, A. Vie, M. Ihme, Characterizing spray flame/vortex interaction: a
17 (12) (20 0 0) 2148–2156. spray spectral diagram for extinction, Combust. Flame 163 (2016) 100–114.
[28] S. Osher, J.A. Sethian, Fronts propagating with curvature-dependent speed: al- [37] M. Nakamura, F. Akamatsu, R. Kurose, M. Katsuki, Combustion mechanism of
gorithms based on hamilton-jacobi formulations, J. Comput. Phys. 79 (1) (1988) liquid fuel spray in a gaseous flame, Phys. Fluids 17 (12) (2005) 123–301.
12–49. [38] H. Nomura, M. Hayasaki, Y. Ujiie, Effects of fine fuel droplets on a laminar
[29] P. Perona, J. Malik, Scale-space and edge detection using anisotropic diffusion, flame stabilized in a partially prevaporized spray stream, Proc. Combust. Inst.
IEEE Trans. Pattern Anal. Mach. Intell. 12 (7) (1990) 629–639. 31 (2) (2007) 2265–2272.
[30] G. Hartung, J. Hult, R. Balachandran, M.R. Mackley, C.F. Kaminski, Flame front [39] X. Mercier, M. Orain, F. Grisch, Investigation of droplet combustion in strained
tracking in turbulent lean premixed flames using stereo PIV and time-se- counterflow diffusion flames using planar laser-induced fluorescence, Appl.
quenced planar LIF of OH, Appl. Phys. B 96 (4) (2009) 843–862. Phys. B 88 (1) (2007) 151–160.
[31] P. Xavier, A. Vandel, G. Godard, B. Renou, F. Grisch, G. Cabot, M.A. Boukhalfa, [40] D.H. Wacks, N. Chakraborty, Flame structure and propagation in turbulent
M. Cazalens, Investigation of combustion dynamics in a cavity-based combus- flame–droplet interaction: a direct numerical simulation analysis, Flow, Turbul.
tor with high-speed laser diagnostics, Exp. Fluids 57 (4) (2016) 50. Combust. 96 (4) (2016) 1053–1081.
[32] J. Hult, U. Meier, W. Meier, A. Harvey, C.F. Kaminski, Experimental analysis of [41] D.H. Wacks, N. Chakraborty, E. Mastorakos, Statistical analysis of turbulent
local flame extinction in a turbulent jet diffusion flame by high repetition 2-D flame–droplet interaction: a direct numerical simulation study, Flow, Turbul.
laser techniques and multi-scalar measurements, Proc. Combust. Inst. 30 (1) Combust. 96 (2) (2015) 573–607.
(2005) 701–709.

You might also like