Characterization of the kerosene spray in a swirl

Transcription

Characterization of the kerosene spray in a swirl
16th Int Symp on Applications of Laser Techniques to Fluid Mechanics
Lisbon, Portugal, 09-12 July, 2012
Characterization of the kerosene spray
in a swirl-stabilized flame
Nicolas Fdida1,*, Christophe Brossard1, Axel Vincent-Randonnier1,
Daniel Gaffié1
1: Onera - The French Aerospace Lab, F-91761 Palaiseau, France
* correspondent author: [email protected]
Abstract Combustion processes involved in aeronautic combustors greatly depend on the turbulent flow
structure, the characteristics of the liquid fuel spray and its atomization. Simulation of these physical and
chemical processes uses complex numerical tools which lean on experimental data. The drop-size of the
spray is an important feature which impacts directly on the computation quality. The aim of this study is to
obtain a large database on the liquid fuel spray characteristics for a swirl-stabilized flame, under conditions
close to industrial combustors. A Phase Doppler Interferometer (PDI) measured the droplet size and velocity
distributions for the steady spray flame. Spatial distributions of these data were compared for two operating
conditions. A recirculation zone was evidenced from PDI results and the swirl motion of the flame was
characterized. The transient ignition phase was investigated, even if the main goal was to study the steady
phase of combustion. Results showed similar mean droplet sizes and velocities in both phases. The PDI
processes the laser scattering signals from droplets passing through a measurement volume with a high
frequency sampling, which is not constant in time. To perform a Fourier transform on droplet time arrival, a
resampling processing is required. Such a sampling algorithm, developed at ONERA, was performed for
several measurement locations. Frequencies close to 1 kHz were evidenced, with values depending on the
measurement location in the flame and on the operating conditions. To explain these frequencies, possible
causes, such as chamber acoustics or swirl vortex breakdown, are examined.
1. Introduction
The combustion process in aeronautic combustors greatly depends on the turbulent flow structure,
the characteristics of the liquid fuel spray and its atomization. To obtain meaningful and reliable
numerical computations of combustion processes, the physical input parameters of the operating
conditions have to be well known. The drop-size distribution of a spray is an important input
parameter because it has a direct impact on numerical results quality. The aim of this study is to
obtain a large database on the liquid fuel spray characteristics for a swirl-stabilized flame, under
conditions close to industrial combustors. For this purpose, a Phase Doppler Interferometer (PDI) is
a useful tool to obtain at the same time the size and velocities of the burning kerosene droplets. In
this paper, spatial distributions of these quantities for the steady spray flame, with comparisons
between both operating points, are presented. A recirculation zone was evidenced from PDI results
and the swirl motion of the flame was characterized. Even if the main goal of the study was to
characterize the steady phase of combustion, particular attention was paid to the transient ignition
phase, which is rarely addressed in the literature. In order to characterize spray and combustion
instabilities experimentally, two different types of signals were analyzed and compared: high-speed
laser planar images, and PDI data. Indeed, the time of arrival of the PDI data can be post-processed
to reveal pulsations and clustering of spray droplets (Bachalo et al., 1990). In this paper, a fast
Fourier transform was performed to detect cluster arrival frequency and periodicity of cluster
formation. In such a spray flame, droplet clusters can be related to the periodic instabilities of the
flame or pollutant formation.
-1-
16th Int Symp on Applications of Laser Techniques to Fluid Mechanics
Lisbon, Portugal, 09-12 July, 2012
2. Experimental Setup
2.1 EPICTETE test Rig
The EPICTETE swirl-stabilized combustion chamber is 1.4-m long and has a square cross-section
of 100 mm x 100 mm (Brossard et al., 2010), as illustrated in Figure 1
Figure 1. The preheated air flow (mair=80 g/s and air temperature around 430 K) enters the
combustion chamber through an industrial swirl-injector (Turbomeca Makila DLN). The kerosene
(Jet-A1) spray, injected horizontally, along the X axis, by a DELAVAN nozzle, is formed at the tip
(pilot flame configuration) of the swirl injector, which is representative to industrial conditions. The
fuel-air equivalence ratio is ER=0.8. The burned gases exit the test rig through an exhaust. An
adjustable needle can modify the chamber counter-pressure by changing the exit throat cross
section. In this study, two conditions of chamber pressure were investigated: 100 kPa and 200 kPa.
Optical diagnostics were performed through large quartz windows of 260 mm length and 100 mm
height, placed on both sides of the combustion chamber. Measurements were performed under
reacting conditions, within the stabilized period of combustion. Each fire test lasted 5 to 15 minutes.
choked
nozzle
Turbomeca swirl
injector
window
throttling
plug
exhaust
combustion chamber
Figure 1 : View of the EPICTETE test rig
2.2 Optical diagnostics
High-Speed Laser Planar (HSLP) visualizations were performed in reacting flow conditions in
order to define the area where the liquid phase is present. A Quantronix Darwin-Dual laser system
was used to generate laser pulses shorter than 270 ns, at 527 nm and 4 kHz repetition rate. A high
speed camera (Photron APX-RS3000) recorded the Mie scattering signal from the spray droplets
(see background image in Figure 2).
A Phase Doppler Interferometer (PDI) from ARTIUM Inc., was used to characterize the kerosene
droplet size and velocity distribution in the reacting flow, under steady operating conditions. A
solid state laser system delivers green (λg=532 nm) and blue beams (λb=491.5 nm). The focal length
of the emitter and receiver lenses was 500 mm. The off-axis angle of the receiver was set to 40° due
to mechanical constraints around the test rig. The refractive index n of the kerosene droplets is
temperature dependent, thus a value of n=1.4 was chosen between the preheated air flow conditions
and the boiling state of the kerosene JP5 (Lefebvre, 1989), through the Eykman relationship
(Pitcher et al., 1990). According to the PDI manufacturer, a temperature range from ambient to the
boiling point results in an uncertainty of less than +/- 3% on the deduced droplet diameter.
Based upon the optical setup, represented in Figure 3 the PDI could measure droplet diameters
within the range 1.3 µm<D<192.8 µm, through the green channel. The number of validated droplet
diameter measurements in each acquisition sample was at least 5,000, except for measurement
-2-
16th Int Symp on Applications of Laser Techniques to Fluid Mechanics
Lisbon, Portugal, 09-12 July, 2012
locations where the droplet density was too low.
The PDI provides two components of the droplet velocity (horizontal Vx and vertical velocity Vy,
respectively through blue and green channel). The PDI instrument parameters were self-optimized
to the velocity range based upon a 300 droplets sample analyzed before a measurement. The largest
velocity range in our acquisitions was -176 m/s < Vx < 528 m/s and -166 m/s<Vy<166 m/s for
horizontal and vertical velocity, respectively.
emitter
Z
Y
X
Vx
O
40°
Vy
receiver
Figure 2: locations of the PDI measurement points in
the vertical plane O,X,Y. Background : instantaneous
HSLP image of the spray.
Figure 3: Top view of the PDI optical
setup in the horizontal plane O,X,Z.
Definition of PDI measured velocities.
2.3 Location of measurement points
PDI measurements can be performed only where droplets are present. The two-phase flow area to
be investigated was defined based upon time-averaged HSLP visualizations. An example of
instantaneous HSLP image is presented in the background in Figure 2 with an inverted gray color
scale. Measurement locations, illustrated by white dots in Figure 2, are located in three horizontal
planes Y=0 and Y=+/- 15 mm. In these planes, transverse profiles were made along the Z axis, from
X=5 mm to X=20 mm, with a 5 mm step. Diameter and velocity measurements near the nozzle exit
(X=1.5mm and X=3mm) were also performed with the green channel only, thanks to the off-axis
angle of the receiver.
3. Results
3.1 Droplet size measurements
Two examples of probability density functions (pdf) are presented in Figure 4.a and b. They are
based on the droplet size distribution by number and representative of the kerosene spray under
reacting conditions, at P=200 kPa and ER=0.8. The corresponding measurement locations are along
-3-
16th Int Symp on Applications of Laser Techniques to Fluid Mechanics
Lisbon, Portugal, 09-12 July, 2012
the injection axis, at X=12.5 mm, Y=Z=0 (Figure 4.a) and on the swirl cone envelope, at
X=12.5mm, Y=-15mm, Z=0 (Figure 4.b). The pdf profile is fairly Gaussian in both cases, centered
on 13µm (Figure 4.a) and on 19µm (Figure 4.b). The droplet density is the largest in the swirl cone
envelope, about twenty times larger (2764 droplets/s, see Figure 4.b), than along the injection axis
(128 droplets/s, see Figure 4.a).
A few very large droplets (around 10 in the acquisition samples of size N, represented in Figure 4)
were observed in the diameter range (50<D<200µm). Due to their low number, they are not visible
in the droplet size distribution by number, but are responsible of a high D32 compared to D10.
a) Injection axis : X=12.5mm, Y=Z=0,
P=200kPa
b) Swirl cone : X=12.5mm, Y=-15mm, Z=0,
P=200kPa
Figure 4 : droplet size distributions of the spray, reacting case, P=200 kPa.
The spatial distribution of the Sauter Mean Diameter (D32) is presented in Figure 5, for both
operating conditions investigated (P=100 kPa left side and P=200 kPa on right side, same scale). In
the background, corresponding time-averaged HSLP images are shown. The swirl cone angle,
estimated through averaged HSLP images, is larger for the 200 kPa case, because the counterpressure results in a more compact flame. A general trend for Figure 5 is an increase of the droplet
size radially towards the outer region of the spray. The smallest droplets are in the swirl core region,
near the injection axis, and the largest droplets are on the outer borders of the spray. This spatial
distribution is probably due to the presence of a recirculation zone which brings the unburned
droplets back in the swirl core, increasing the residence time and droplets evaporation in this region
(see Pascaud et al., 2005). Figure 5.b also shows that the droplet size decreases with the distance
from the injector.
-4-
16th Int Symp on Applications of Laser Techniques to Fluid Mechanics
Lisbon, Portugal, 09-12 July, 2012
X axis (mm)
X axis (mm)
-5 0
-5 0
5 10 15 20 25 30 35 40 45 50
5 10 15 20 25 30 35 40 45 50
-50
-50
-45
-45
50µm
-40
-40
-35
-35
-30
-30
-25
-25
-20
-20
50°
-15
35°
-15
50µm
-10
-10
-5
-5
Z axis
0
(mm)
Z axis
0
(mm)
5
5
10
10
15
15
20
20
25
25
30
30
35
35
40
40
45
45
50
50
a) P=100kPa.
b) P=200kPa.
Figure 5 : Spatial distribution of D32 in the horizontal plane Y=0. Dotted lines indicate the swirl
cone angle estimated from time averaged HSLP images (background image).
3.2 Velocity measurements
Axial velocity distributions (not shown here) are different in the recirculation zone than in the spray
cone region. Indeed, in the central zone, a population of droplets with negative velocities (Vx≈50 m/s) is mixed with droplets of high axial velocity (Vx≈150 m/s). Outside from the injection axis
region (for Z≤±5mm), only positive velocities are observed.
Figure 6 shows the spatial distribution of axial velocities Vx in the horizontal plane Y=0, for the
operating condition P=200 kPa. The spatial variations of Vx are stronger radially than axially, as
observed for the mean droplet size in Figure 5. This trend is similar for the other operating
condition P=100 kPa (not shown here). Negative axial velocities are observed in the region close to
the injection axis, corresponding to 5≤ X ≤20 mm for P= 200 kPa (Figure 6) and 10≤ X ≤30 mm for
P= 100 kPa (not shown here). For such a swirling flow, the transition from negative to positive
axial velocities defines the stagnation point of the vortex breakdown, according to Lucca-Negro and
O’Doherty, 2001. The axial velocity reaches a maximum for measurement points in the swirl cone
envelope, and a minimum in the injection axis. The locations of the maxima are in agreement with
-5-
16th Int Symp on Applications of Laser Techniques to Fluid Mechanics
Lisbon, Portugal, 09-12 July, 2012
Z axis(mm)
the dotted lines indicating the location of the swirl half cone angle, deduced from time-averaged
HSLP images (see background images in Figure 5.b). We also noticed that the half-cone angle was
slightly smaller in the region Z > 0. No particular size/velocity correlation was revealed from the
PDI data analysis, in any location of the spray, neither on Vx nor on Vy.
50
45
40
35
30
25
20
15
Vertical Velocity Vy
(m/s)
100
80
60
40
20
10
5
0
0 -5 -20
-40
-60
-80 -100
-10
-15
-20
-25
-30
-35
X=1.5mm
X=10mm
X=20mm
-40
-45
-50
Figure 6 : Spatial distribution of mean axial
velocity Vx in the horizontal plane Y=0,
P=200 kPa. Dotted lines indicate the 50° half
cone angle of the swirl.
Figure 7 : Radial profiles of the mean vertical
velocity Vy (m/s) in the plane Y=0, P=200 kPa.
Three radial profiles of the vertical velocity Vy are presented in Figure 7, for P=200 kPa. The swirl
motion of the flow is clockwise, which is in agreement with the orientation of the nozzle vanes (see
picture in Figure 7). The vertical velocity profiles from X=1.5 mm to X=20 mm show a global
conservation of the rotating motion of the swirl. The same general trends are observed for both
operating conditions (P=100 kPa and P=200 kPa), indeed the absolute maxima of the radial profiles
decrease as X increases, and the position of these maxima is shifted outwards radially, thus
indicating a small expansion of the swirl cone. The maxima of vertical velocity recorded in the
radial profiles in the horizontal mid-section plane Y=0 correspond to the swirl tangential velocities.
Therefore, the turnover time of the swirling flow could be estimated at τswirl ≈1.35 ms based upon
the radial locations of the maxima in the profile X=1.5mm close to the nozzle. By using the Vx and
Vy radial profiles at the nozzle exit, the value of the swirl number S can be calculated. For the
P=200 kPa operating condition, a value of S≈0.65 was obtained, which is consistent with the results
from Meier et al., 2007, obtained with the same injector but for methane combustion.
The velocity profiles also allowed to characterize the axisymmetry of the spray. The spray was
found to be fairly axisymmetric with respect to the injection axis.
-6-
16th Int Symp on Applications of Laser Techniques to Fluid Mechanics
Lisbon, Portugal, 09-12 July, 2012
3.4 Study of the ignition phase
The whole measurement campaign was performed through several fire tests, which naturally began
by an ignition phase. As the PDI sampling rate is in the MHz range, it is possible to focus on data
acquired during the ignition phase. In Figure 8, the temporal evolution of the diameter measurement
of the first droplets, during the ignition phase, is presented for the measurement location X=20mm,
Y=0, Z=0, P=100 kPa. The flame ignition was performed at ER≈0.9, slightly above the nominal
condition. Following two bursts of droplets (detected at t=11.5 s and t=13.5 s), the combustion
began at t=14.2 s. The droplet diameters of the first detected droplets are close to the one in the
combustion phase. During the ignition, as well as during the beginning of the combustion, a few
droplets of large diameter (D≈180µm) were observed. The values of the droplet size and velocities
(not shown here) remain unchanged from the ignition phase to the steady phase of combustion.
Figure 8: temporal evolution of droplets diameter in the ignition phase. X=20mm, Y=0, Z=0,
P=100 kPa
3.5 Frequency analysis of instabilities
Instantaneous HSLP images, recorded at 4 kHz, revealed high spatial and temporal variations. A
Fast Fourier Transform (FFT) was applied on the instantaneous mean gray intensity levels, defined
by spatially averaging the intensity over the whole image. This analysis, performed over the steady
phase, revealed a main instability peak at a frequency around 1kHz: 0.840 kHz for ER=0.8,
P=100 kPa and 1.126 kHz for ER=0.8, P=200 kPa. For the P=100 kPa case, the first harmonic
frequency is also visible (see Figure 9, top, dotted line). For the P=200 kPa case, an additional
characteristic frequency of ~2000 Hz is also revealed by the FFT (see Figure 9, bottom, dotted line).
However the validity of this 2000 Hz frequency peak deduced from the analysis of HSLP images is
questionable, as it is only half of the 4 kHz frame rate of the images.
The FFT analysis was also conducted for the P=100 kPa case for various increasing equivalence
ratios in the range 0.7<ER<1.2, and led to an increase of the main frequency from 820 Hz to
885 Hz. Moreover, for ER=0.8, the main frequency increased from 830 Hz to 1130 Hz when the
-7-
16th Int Symp on Applications of Laser Techniques to Fluid Mechanics
Lisbon, Portugal, 09-12 July, 2012
chamber pressure was increased from 150 to 200 kPa. Therefore, these instabilities frequencies
clearly depend upon the flame operating conditions. In an attempt to discriminate the frequency
origin, the FFT analysis was also conducted locally, by spatially averaging the intensity over
different regions of interest of 10x10 mm² in HSLP images, rather than in the whole image. Three
regions were selected, centered on X=12.5 mm and Y=0 and Y=±15 mm. However, these ‘local’
FFTs (not shown here) led to the same results.
Figure 9: HSLP and PDI Fast Fourier Transform spectra for P=100 kPa (top) and 200 kPa
(bottom), ER=0.8
The high sampling rate of PDI data (MHz range) was utilized to obtain the frequency content of the
spray through FFT analysis. Different regions of the spray were investigated: in the swirl core
region and in the swirl envelope. Because the PDI sampling is random, a re-sampling is required
prior to FFT analysis. The re-sampling Refined Sample and Hold algorithm by Nobach et al., 1998,
was applied through the ASSA software, developed at ONERA by Micheli et al., 2006. A 4 kHz
cut-off frequency limit was applied in the FFT computations. We applied the FFT on arrival times
of axial velocity data Vx, corresponding to the green channel, which always exhibited a better data
rate. Figure 9shows the FFT spectra obtained for three different PDI measurement points for each
case investigated. These measurement points were located in the vertical mid-section plane (Z=0) at
Y=0, +15 and -15mm; due to different swirl angles (see Figure 5), two different axial locations
were chosen, X=20 mm, for the P=100 kPa case and at X=12.5 mm for the P=200 kPa case. No
frequency peak was detected in the 2200-4000 Hz frequency range, therefore this range is not
presented in Figure 9. PDI revealed the same frequencies as HSLP images (f≈1150 Hz for
P=200 kPa and 850Hz for P=100 kPa), except for the measurement point located on the injection
axis X=12.5mm, Z=0 and Y=0. From the PDI data analysis, the 2000 Hz frequency was only
detected in the swirl envelope (X=12.5mm, Z=0 and Y=±15 mm, see Figure 9, and X=12.5mm,
-8-
16th Int Symp on Applications of Laser Techniques to Fluid Mechanics
Lisbon, Portugal, 09-12 July, 2012
Y=0 and Z=±15 mm, not shown here) and not on the injection axis (X=12.5mm, Z=0, Y=0).
Therefore, this 2 kHz frequency could be specific to the swirl region. PDI FFT analysis also
revealed frequencies in the range 800-900 Hz (blue, red and green solid line peaks, Figure 9,
bottom) for the P=200 kPa case, that are not detected in the FFT analysis of the HSLP images.
In the operating conditions investigated in this paper, the swirling flow generated by the nozzle is
characterized by a swirl number S of ≈0.65 (see section 3.2). According to Lucca-Negro and
O’Doherty, 2001, such a high swirl number can induce an axial vortex breakdown, associated with
a Precessing Vortex Core (PVC). Pascaud et al., 2005, estimated the PVC characteristic frequency
through the relationship f=1/τswirl. In the present experiment, the turnover time was found to be
τswirl≈1.35 ms for the P=200 kPa case investigated in this paper (see section 3.2). Therefore, it
provides an estimation of the characteristic frequency of f≈743 Hz. On one side, this value is close
to frequencies in the range 800-900 Hz detected in the FFT analysis of the PDI signals (blue, red
and green solid line peaks, Figure 9, bottom), but not in the FFT analysis of the HSLP images. This
could be explained by a much stronger sensitivity of the PDI to PVC motion due to the very high
spatial resolution of this technique (very small measurement volume). On the other side, Poinsot
and Selle, 2005 characterized the PVC numerically for non-reacting and reacting swirling flow
conditions, and indicated inhibition of the PVC by the reacting flow. Thus the aerodynamic origin
of these frequencies has not been clearly identified yet.
The instabilities are not attributed to a coupling with the acoustic modes of the chamber. Indeed,
considering the operating conditions at P=100 kPa and the combustion chamber dimensions, the
first resonant acoustic modes are 408 Hz (first longitudinal mode), 4487 Hz (first transverse modes)
and even higher for tangential modes. The value of 408 Hz does not match the frequencies
evidenced by PDI or HSLP images. In order to fully identify the phenomena involved with the
instabilities evidenced in this paper, additional experimental data are required, such as dynamic
pressure sensors placed at different locations in the chamber walls, and high-speed OH*
chemiluminescence images.
4. Conclusion
The droplet size and velocity distributions in a swirl-stabilized kerosene/air combustor
representative of aeronautical combustors were investigated by Phase Doppler Interferometry. Two
different pressure conditions, P=100 kPa and 200 kPa, were studied. Spatial profiles of droplet size
and 2-components velocity measurements were analyzed in three horizontal planes and close to the
nozzle exit. Droplets were found to be slightly larger in the swirl cone envelope than along the
injection axis, due to the presence of a central recirculation zone, evidenced by negative axial
velocities measured along the injection axis. The axial and vertical velocities reach a relative
maximum for measurement points in the swirl cone envelope and a minimum in the injection axis
region. The locations of the maxima are in agreement with the location of the swirling spray cone
visible in the averaged planar laser sheet image of the spray. Comparisons of the results obtained
for the two operating conditions investigated showed similar trends for the spatial profiles of
velocities and diameters. However, slightly smaller droplet sizes, as well as and a smaller half-cone
angle of the swirl, were observed for the 100 kPa case.
The high frequency sampling of PDI data allowed to study the ignition phase and the frequency
content of instabilities. This frequency content was found to be consistent with the one deduced
from high-speed laser images. Both techniques showed frequencies in the kHz range, the exact
values of which depended upon the experimental conditions. In order to fully identify the
phenomena involved with the instabilities evidenced in this paper, additional experimental data
would be required from a future work, such as dynamic pressure sensors placed at different
locations in the chamber walls, and high-speed OH* chemiluminescence images.
-9-
16th Int Symp on Applications of Laser Techniques to Fluid Mechanics
Lisbon, Portugal, 09-12 July, 2012
5. References
Bachalo W., Rudoff, C.R. and Sankar, S.V. (1990), Time-resolved measurements of spray drop size
and velocity, In: Hirleman, Bachalo, Felton Eds., Liquid particle size measurement techniques, 2nd
volume, American Society for Testing and Materials, pp.209-224.
Brossard C., Vincent-Randonnier A., Barat M. and Gicquel P. (2010), Caractérisation par PIV haute
cadence d’une combustion stabilisée par injection tourbillonnaire, CFTL 2010, Vandoeuvre-lèsNancy.
Lefebvre, A.H. (1989), Atomization and Sprays, Hemisphere publishing Corporation.
Lucca-Negro and O’Doherty (2001), Vortex breakdown: a review, Prog. Energy Comb. Sci. Vol.27,
pp.431-48.
Meier W., Weigland P., Duan X.R., Giezendanner-Thoben R. (2007), Detailed characterization of
the dynamics of thermoacoustic pulsations in a lean premixed swirl flame, Combustion and Flame,
vol.150, pp.2-26.
Micheli F., Lavieille M., Millan P. (2006), ASSA, un outil de référence pour le traitement du signal
en vélocimétrie laser, 10ème Congrès Francophone de Techniques Laser (CFTL), Toulouse, France.
Nobach H., Müller E.and Tropea C. (1998), Correlation estimator for two-channel, non-coincidence
laser-doppler-anemometer, 9th International Symposium on Applications of Laser Techniques to
Fluid Mechanics, Lisbon, Portugal.
Pascaud S., Boileau M., Martinez L., Cuenot B. and Poinsot T. (2005), Large Eddy Simulation of
turbulent spray combustion in aeronautical gas turbines, ECCOMAS Thematic Conference on
computational combustion, Lisbon, Portugal.
Pitcher, G., Wigley, G., and Saffman, M. (1990), Sensitivity of dropsize measurements by phase
Doppler anemometry to refractive index changes in combusting fuel sprays, Fifth International
Symposium on Applications of Laser Techniques to Fluid Mechanics, Lisbon, Portugal, pp. 227247.
Poinsot T.and Selle L. (2005), LES and acoustic of analysis of combustion instabilities in gas
turbines, In PLENARY LECTURE - ECCOMAS Computational Combustion Symposium,
Lisbonne, Portugal.
Acknowledgment
This work was performed with financial support from the French DGA (Délégation Générale de
l’Armement) and ONERA.
- 10 -