You are on page 1of 37

This article was downloaded by: [University of Tennessee, Knoxville]

On: 12 May 2013, At: 04:27


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Combustion Science and Technology


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/gcst20

An Experimental and Numerical Study of Kerosine


Spray Evaporation in a Premix Duct for Gas Turbine
Combustors at High Pressure
a a a
M. BRANDT , M. RACHNER & G. SCHMITZ
a
German Aerospace Research Center (DLR), Institute for Propulsion Technology, Koin,
D-51170, Germany
Published online: 25 Apr 2007.

To cite this article: M. BRANDT , M. RACHNER & G. SCHMITZ (1998): An Experimental and Numerical Study of Kerosine Spray
Evaporation in a Premix Duct for Gas Turbine Combustors at High Pressure, Combustion Science and Technology, 138:1-6,
313-348

To link to this article: http://dx.doi.org/10.1080/00102209808952074

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to
anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents
will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should
be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims,
proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in
connection with or arising out of the use of this material.
CombUS1. Sci. and Tech., 1998. Vol. 138, pp. 313-348 © 1998OPA (Overseas Publishers Association)N.V.
Reprints available directlyfrom the publisher Published by license under
Photocopyingpermitted by licenseonly the Gordon and Breach Science
Publishers imprint.
Printed in Malaysia.
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

An Experimental and Numerical Study


of Kerosine Spray Evaporation in a Premix
Duct for GasTurbine Combustors
at High Pressure
M. BRANDT, M. RACHNER* and G. SCHMITZ

German Aerospace Research Center (DLR), Institute for Propulsion Technology,


D-51170 Koln, Germany

(Received 27 Apri/199B; In final form 20 July 1998)

The evaporation and mixing of kerosine emerging from a flat prefilming airblast atomizer was
studied experimentally and numerically in an optical accessible, straight rectangular duct at
conditions relevant for lean premixed and prevaporized combustion. Liquid phase properties
were measured by Phase-Doppler anemometry and fuel vapour concentrations were determined
by an infrared light extinction technique. Computations were based on the Lagrangian particle
tracking technique, and captured the spray features sufficiently well with and without taking
into account the spray feedback on the gas field. A degree of evaporation of95% was measured
after IOOmm for 9 bar air pressure, 750K air temperature and 120m/s air velocity. No
autoignition of the fuel occurred. Parametric variations of pressure, temperature and velocity of
the air flow, as well as of the initial temperature of the spray and of fuel loading were conducted.
A strong influence of the initial fuel temperature on evaporation was found. The air pressure
had mainly an indirect but strong effect through atomization quality.

Keywords: Spray evaporation; high pressure; gas turbine combustion chamber; prefilming
airblast atomizer; premix duct; kerosine

1. INTRODUCTION

The NO x emissions from gas turbine combustors are known to increase


steadily with increasing pressure ratios and combustor inlet temperatures.
One combustion concept to reduce NO x emissions is lean premixed and
prevaporized combustion (LPP). This concept relies on a lean, homo-

·Corresponding author. e-mail: michael.rachner@dlr.de

313
314 M. BRANDT et al.

geneous fuel/air mixture burning at low temperature and therefore low


formation rates of thermal NO x . For gaseous fuels in stationary gas turbines
lean premixed combustion has been demonstrated to cause a 90% decrease
of nitric oxide emissions compared to standard diffusion burners (Jansen et
al., 1991). Drawbacks of this concept are its narrow operating range, the
proneness to combustor noise and -rnainly for liquid fuels- the risk of
flashback and autoignition. The autoignition times of aircraft-type fuels are
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

about a few milliseconds at typical combustor inlet conditions (Spadaccini


and TeVelde, 1982) and therefore impose a limitation on the preparation of
liquid fuels for lean, premixed and prevaporized combustion. The fuel has to
be completely evaporated and should be mixed homogeneously with the air
in the premix duct before the time of autoignition elapses.
A typical air velocities (~ 100m/s) and air densities in premix ducts, a very
fine spray can be produced by airblast atomizers. However this result
depends strongly on liquid and air properties, which in turn are governed by
the duct operating conditions. Droplet evaporation and dispersion are
influenced by the initial conditions of the spray produced and depend on the
duct operating conditions too. Therefore the investigation of such a
premixing passage should be done at realistic air pressures, temperatures
and velocities and with a typical fuel.
In the present work, the evaporation rate, dropsize distribution, droplet
velocities and -temperatures, as well as liquid and vapour concentrations of
evaporating fuel sprays in the premix duct are presented. A kerosine Jet A-I
spray produced by a flat prefilming airblast atomizer has been investigated
at conditions relevant to a premix duct for aeroengines and high-
performance stationary gas turbines. The experiments were carried out in
an optically accessible generic prevaporizer model that allowed the use of
non-intrusive measurement techniques. Numerical predictions were based
on Lagrangian particle tracking. The comparison of numerical predictions
with the experimental results improved the understanding of the underlying
physics and did separate effects that could not be distinguished by the
experiment alone. The results of such a comparison as well as a parametric
study of the influence of the operating conditions on the initial spray
conditions and the evaporation rate are presented here.

2. EXPERIMENT AND MEASUREMENT TECHNIQUE

2.1. Test Cell


The experiments were carried out in a high pressure, high temperature test
cell designed for three-way optical access to a straight premix duct (Fig. I).
KEROSINE SPRAY EVAPORATION 315

Quartz-duct
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

z z
Cross section L y
Longitudinal section L x
FIGURE I Test cell.

A rectangular quartz-glass duct with a cross-section of 25 x 40 mm and a


maximum observeable streamwise. pathlength of about l80mm was
mounted inside the pressure housing. Air temperatures up to 850 K at. air
velocities of about 120 mls and air pressures up to 15 bars were achieved
with a 520 kW electrical air heater. Inside the rectangular premix duct,
atomizers could be mounted at two axial positions to keep the fuel residence
time (computed by the mean air velocity and the pathlength) below the
autoignition time of the fuel, which depends on the operating conditions.
The quartz glass duct was surrounded by a cooling airflow, to isolate the
pressure housing from the high temperatures. Both main air and cooling air
flow could be controlled separately and left the test cell through a common
throttle. At a given main air temperature, the pressure inside the premix
duct was regulated by a variation of the air mass fluxes and the use of
throttles of different diameter at the duct exit. Volume fluxes, pressures and
temperatures of the air were measured with vortex meters, pressure
transducers and NiNiCr thermocouples, respectively. Due to a slow drift
in air supply pressure, moderate deviations 6.TAir = ±2.5K, 6.uAir =
±2m/s, 6.PAir = ±0.15 bar from the nominal operating conditions in the
test cell occurred. These variations of the air pressure in the test cell also
caused fuel mass flow changes of ± 0.02 gis, i.e., 2% for the baseline case
(Chapter 4. I).

2.2. Atomizers

The speed of evaporation of a fuel spray is strongly influenced by its initial


dropsize distribution. To achieve high evaporation rates in a given time,
316 M. BRANDT et 01.

which is limited by the autoignition of the fuel, the smallest possible dropsize
distribution has to be produced by the atomizer. Hence, in a previous
experimental study fuel injectors were tested for their suitability for use in
premix ducts at engine operating conditions (Brandt et al., 1994).
Consequently, the atomizer producing the lowest dropsize diameters, a flat
prefilming airblast atomizer (Fig. 2) was chosen for this investigation.
For this type of atomizer the fuel is spread out on the surface of the
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

atomizer, where it forms a thin film, which is atomized by the high-


momentum air stream. Downstream (x > 0) of the atomizer edge the 'spray
is then accelerated to the air velocity. All studies except the parameter
variation of the atomizer fuel loading were conducted with the small
prefilming atomizer of 8 mm width and a slit height of 0.1 mm. Its prefilming
plane coincides with the horizontal symmetry plane (x - y plane) of
the duct. The origin of the coordinate system is located in the middle of
the atomizer edge, which is in the center of the cross-section of the duct.
The fuel loading ratio I (= rue! mass flux per atomizer width) was chosen so
as to correspond to fuel/air ratios in premix ducts.
The orientation of this atomizer gave a good optical access perpendicular
to the prefilming plane for the light extinction technique, which yields

Insulation

21
L, Topview

2 ~:=:J~rF~~~~~ I 0.1 (0.3) z

4
NiNiCr Thermocouple
L, Sideview

FIGURE 2 The small, flat prefilming airblast atomizer (dimensions in brackets refer to the
larger atomizer).
KEROSINE SPRA Y EVAPORATION 317

integral values of extinction along the horizontal y-axis, i.e., along the
direction of the atomizer edge. For the fuel loading variation another
atomizer with a larger width of 30 mm and a larger slit height of 0.3 mm
(values in brackets in Fig. 2) was used. It had to be mounted vertically in the
duct (prefilming plane was now the x - z plane at y = 0). Both atomizers
were made of the same stainless steel, with a comparable wall thickness and
about the same fuel path length in the duct.
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

The evaluation of the experimental results revealed a strong influence of


the initial fuel temperature on atomization and evaporation. Therefore,
after having finished most of the experiments, a NiNiCr thermocouple (bead
diameter 0.1 mm) was embedded in the surface of the small atomizer. As a
protection against thermal conduction from the atomizer housing, it
was isolated by a ceramic bond. More than 70% of the bead volume was
embedded in the isolator, so that the film flow on the atomizer surface
was not disturbed by the bead.

2.3. The PDA- and Light Extinction Setup

The setup of measurement instruments is shown in Figure 3.

I. IR-Detector (pbS-Detector)
(400 flm Aperture Diameter)
2. PDA Receiving Optics
3. Lens (CaF, f= 50 mm)
4. Photodetector (Si Diode)
(150 urn Aperture Diameter)
5. Lens (f= 100 mm)
6. Color Splitters
7. Test Cell
8. Chopper
9. Mirrors
10. Lens (f= 300 mm)
11. Lens (CaF, f=400 mm)
12. PDA Transmitting Optics
13. HeNe Laser (633 nm)
14. HeNe Laser (3.39 urn)

FIGURE 3 Setup of PDA and light extinction technique.


318 M. BRANDT et al.

2.3.1. The Phase Doppler System

Droplet sizes and -velocities as well as liquid volume fluxes were measured
by a two component Phase Doppler Anemometer (PDA) with a covariance
processor, see DANTEC (1992). The Gaussian beam diameter in the
measurement volume was about 70 11m, the laser-power was typically 25 mW
per beam. Drop sizes assuming spherical particles were measured in first
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

order refraction. The light was polarized parallel to the scattering plane so
that at a scattering angle equal to the Brewster's angle of the fuel (~69° for
kerosine at ambient temperature), light reflection is completely absent.
However the scattering angle was limited by the optical access of the test cell
to 52°, which led to an intensity ratio of refracted to reflected light of 180.
For this setup, velocity and dropsize measurement errors due to
alignment uncertainties were below 1%. Neglecting refractive index
gradients, dropsizing errors due to the change of the refractive index of
kerosine for droplets being heated up from ambient temperature to their
wet-bulb temperature were about 3.5%. This error was computed with the
decrease of the refractive index of kerosine with temperature, which was
measured by Fitzky and Bohn (1995). In the results presented here, a
constant refractive index n = 1.46 of kerosine at ambient temperature was
used, therefore the droplet diameter are tendentially measured too large.
Dropsize measurements up to x = 30 mm from the atomizer edge were
aggravated by the atomization (droplet oscillations) and acceleration of the
liquid. Accelerated droplets might be elliptical rather than spherical. A
maximum dropsizing error might be estimated as follows: the Bond number
(Bo := ppapD~/a, where ap is the droplet acceleration) is a measure of the
acceleration force versus the surface tension force of the droplet. At
x = 30mm downstream from the atomizer, Bond numbers of 2.9 were
calculated for droplets of 261lm diameter, which represents roughly an
upper limit of the dropsize distribution of the baseline case (see Fig. 5).
According to Clift et al. (1978) this Bond number leads to unspherical
droplets with a ratio of the minor to major axis of about 0.7. The dropsize
measured by the PDA depends on the orientation of the unspherical
droplets with respect to the optical axis. Assuming that the PDA measures
randomized distributed diameters between the minor axis and the major axis
of rotational ellipsoids, we find that compared to a sphere that contains the
same volume as the ellipsoids, the mean value of the dropsize is measured
about 7% too large.
Evaporation rates of the fuel spray were obtained from liquid volume flux
measurements, which require accurate determination of the size of the
KEROSINE SPRA Y EVAPORATION 319

effective measurement volume of each particle size class. A sketch of the


PDA measurement volume is given in Figure 4. As the elliptical
measurement volume of the laser beam intersection was limited by a
narrow slit in the receiving optics, a postprocessing algorithm was written,
that takes into account that the main flow was in the x-direction. Therefore
the measurement volume of each particle size class was calculated as a
cylinder, the side areas of which were obliquely cut off by the projection of
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

that slit and were crossed by the main flow direction. The algorithm used by
the instrument software requires, that the burst length of each particle
should be limited by the Gaussian beam diameter of the measurement
volume. However here, only a small amount of particles in the vicinity of the
middle of the measurement volume fulfils this condition. This leads to
smaller mean burst lengths of a particle sample, thus to a smaller
measurement cross-section and to higher measured volume fluxes. This
error was accounted for by correcting the measured burstlengths with a
simple formula derived from the geometry of the sending and receiving
optics as shown in Figure 4.
The burst length is the individual pathlength of a particle crossing the
measurement volume, and is approximated by bl = transit time- Ju~ + IV~.
Herein, Up and IVp are the particle velocity components in the x- and z-
direction. The burst lengths bl are measured over the total measurement
volume length L = L 1 + L 2 , as seen by the receiving optics. Assuming, that
all particles of a given size class cross the measurement volume with the
same probability, in section L 1 a mean burst length bl tr ue/2 is measured and

~_---'....-__-++-~__\-_~ I d,
D e-2 +------+ y

De_2 Gaussian beam diameter


d, fringe spacing
s slit width x
x mainflow direction
<p scattering angle
FIGURE 4 Sketch of the PDA measurement volume.
320 M. BRANDT et al.

GO 100
D

50
VEVArJV 0
E ;;g
.
..:!.
40
e...
-
Q)
Q)

E
30
°90% -
.> 0

~
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

Cll >
'6 20 .>
UJ

0 0
S
0
20
10 D D
°10%

0 0
0 50 100 150 200 250
x [mm]
FIGURE 5 Baseline case: streamwise development of characteristic diameters and evapora-
tion rate (symbols = measurement, lines = prediction).

in section (L 2 - L,) a mean value bltrue is measured. Then for the arithmetic
mean value blmeasured over all burst lengths bl the following equation holds:

bl _ L, . (bltrue/2) + (L 2 - LIl . bltrue + L, . (bltrue/2)


measured - LI + L2
L2 . bltrue
= (I)
L, + L2

or:

(2)

Ignoring the slightly elliptical form of the measurement volume we find


L, = De_2(tan'P and L 2 = s(sin 'P. D e - 2 is the Gaussian beam diameter,
which is the radial limit of the laser beam, where the light intensity is
reduced to l(e 2 of the maximum intensity and 'P the scattering angle. That
leads to:

bl _ bl . (De-d tan 'P) + (s/ sin 'P)


true - measured ( /. ) (3)
S SIn 'P

As the PDA demands that only one particle be in the measurement


volume at a time, the algorithm includes a correction of the liquid volume
KEROSINE SPRAY EVAPORATION 321

flux measurements by Poisson statistics (Edwards and Marx, 1992). In this


correction the dead time of the data processor is also included. However, if
particle concentrations were too high (i.e., close to the atomizer edge,
x < 30mm, where concentrations were above 107 particles/crrr' and data
rates of more than 150kHz were measured), the limits of this correction
were exceeded. This can be seen in Figures 5, 7 and 14 at x = 30mm, where
the evaporation rates for T A i r = 750 K and TAir = 650 K, respectively, at
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

p = 9 bar are measured too high. Edwards and Marx (1992) also showed
that dropsize distribution measurements are almost insensitive to particle
rejection rates up to 90%. Therefore dropsize information but no
evaporation rates are presented for upstream locations in Figures 12 and 17.
Test measurements of non-evaporating kerosine sprays in cold air, but at
realistic values of air velocity, densities and fuel flow rate showed that the
errors of the volume flux measurements were within ± 10%, if the particle
concentrations were not too high. With the exception of some evaporation
rates at x = 30 mm, the total error in the volume flux measurements results
from this error of 10% plus the dropsizing error of 3.5% due to refractive
index changes of the kerosine. The droplet volume is proportional to D 3 and
therefore the volume flux is measured up to 10% too large. These two errors
are assumed to be independent from each other, and lead with Gaussian
error propagation to error bars of + 10% and -14% of the.measured liquid
volume fluxes.

2.3.2. The Light Extinction Technique


Hydrocarbon fuels show strong absorption bands in the infrared range close
to a wavelength>' of 3.39 11m, caused by C- H stretching vibrations. In this
study relative vapour concentrations of the evaporating fuel spray were
measured with a simultaneous infrared (>' = 3.39 11m) and visible light
(>' = 633 nm) extinction technique, as shown in Figure 3. The laser beams
were slightly focussed, resulting in a Gaussian beam diameter of 800 11m in
the duct. A detailed description of the setup can be found in Brandt (1995).
The extinction of the infrared laserbeam passing the fuel spray is mainly
caused by absorption by the fuel vapour and the scattering and absorption
by fuel droplets. The amount of fuel drop scattering and absorption in the
infrared range can be estimated from light extinction of the visible laser
beam (Drallmeier, 1994), where fuel absorption can be neglected. This
amount is removed from the infrared extinction measurement, which then is
a measure of the fuel vapour concentration. The solid angle~ of light
322 M. BRANDT et at.

detection were adjusted to the different Mie parameter of the same particle
at different wavelengths. Computations of the Mie-scattering based on
dropsize distributions measured simultaneously by PDA showed that the
amount of light scattered and absorbed by particles in the infrared range can
be measured with an error of ± 10%. However, as kerosine is a
multicomponent fuel, the composition of which depends on its degree of
evaporation, and as the absorption coefficient of the fuel vapour depends on
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

the ambient pressure and temperature, only relative values of the fuel
vapour concentration were obtained with this technique.

3. THE SPRAY CODE

The numerical predictions of the evaporating spray were performed by a


proprietary Lagrangian particle tracking code. For each computational
particle representing a number of droplets, the equations for mass, internal
energy, momentum and location, i.e., eight ordinary differential equations,
were solved by an Adams predictor-corrector scheme with automatical
adaption of order and stepsize (Shampine and Gordon, 1975). The
instantaneous properties of particles passing fixed control areas and
-volumes are registered and finally local, time averaged information about
the liquid phase is obtained from these data.
A dilute spray (i.e., no particle-particle interaction or droplet breakup)
with spherical droplets is considered. In the momentum equations only the
drag term (employing Putnam's CD-formula) needed to be taken into
account, as the density ratio of liquid fuel to air is large. A standard
evaporation model (quasi-steady spherically symmetric Stefan flow around
droplet with Ranz- Marshall convection corrections of Nusselt- and
Sherwood numbers; variable Lewis number; phase equilibrium at droplet
surface) is used. Real gas effects are not accounted for, as the present case is
well below critical pressure. No local temperature gradients are presumed
inside the droplet, which is considered to be 'well stirred'. The temperature
dependence of all liquid and gaseous properties is taken into account, and
properties of the boundary layer are taken at the 1/3-reference state of
temperature and composition. The multicomponent kerosine fuel Jet A-I is
represented as a single-component fuel. Turbulent particle dispersion is
modelled by a spectral dispersion model (Bliimcke et al., 1993). Neither in
the evaporation model nor in the dispersion model were any constants
tuned. A detailed description of the spray model used is given in Brandt et al.
(1991).
KEROSINE SPRAY EVAPORATION 323

With this spray code a parametric study parallel to the experiments in the
premix duct was performed. For reasons of computational economy, a
homogeneous gas field with no influence of the spray on the gas flow was
assumed. For the baseline case, 195900computational particles were tracked,
requiring about 3.5 hours on a R4400-SGI IndigoZ workstation. Addition-
ally, in order to take into account full coupling between spray and gas flow in
the duct, the TRUST-code developed at DLR Schutz et al., 1997), which is
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

based on the 3d instationary KIVA II-code, was employed. Herein the k -E-
turbulence model was used. The spray models in TRUST and in the
proprietary spray code are virtually the same. Spray feedback on the gas phase
in TRUST was done by accumulating spray source terms of mass, energy,
momentum and turbulent kinetic energy arising from passing particles in
finite volume cells of the gas solver (particle-souce-in-cell method).
The spray modelling requires detailed property data, which are provided
in a comprehensive, critical compilation of property formulas for kerosine
Jet A-I by Rachner (1998).

4. RESULTS AND DISCUSSION

4.1. The Baseline Case

Arising from typical inlet conditions of gas turbine combustion chambers, a


basic operating condition for our investigations in the premix duct was
chosen and is henceforward referred to as the baseline case. It is
characterized by a duct inlet pressure of p = 9 bar, an air inlet temperature
of T A i r = 750 K and a bulk air velocity (= air volume flow/duct cross-
sectional area) UAir = 120tu]«, This corresponds to a ratio of dynamic to
static pressure of 3.3% Fuel mass flow of the kerosine Jet A-I was I gis,
corresponding to a fuel loading (per length of atomizer edge) I = 125g/s/m
of the small flat airblast atomizer used. The results of measurements and
predictions to be presented in the Chapter 4.1 refer to the baseline case.
The PDA measurements of dropsizes, droplet velocities and particle fluxes
were performed in four planes x = 30, 60, 100 and 150mm downstream of
the atomizer edge position x = O. Measurements upstream of x = 30mm
were not possible due to the dense spray limitations (non spherical droplets
and more than one single particle in the measurement volume) of the PDA-
measurement technique. In every measurement plane 50 to 80 measurement
points were located each being limited by a maximum of 20000 events or a
measurement time of 10 sec.
324 M. BRANDT et al.

As will be seen later (Fig. 10), the numerical prediction of the baseline
case and the parameter study could be performed with good accuracy in a
gasfield without gradients and without spray feedback. This computation
considers a spatially homogeneous gas field for the spray with p = 9 bar,
UAir = 125m/s (> 120m/s, as the spray moves near the maximum of the flat
turbulent streamwise gas velocity profile in the duct cross-section),
VAir = WAir = 0, T A i r = 743 K « 750 K in order to take account of the
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

heat losses in the duct), UrmsAir = 7 ml», VrmsAir = Wrms Air = 5 mls (typical
values from measurement), and a turbulent macro length scale
L tu r b = 0.005 m. Spray initial conditions are crucial and are discussed below
(Chapter 4.1.1).
The most important quantity for characterization of the evaporation
process in the duct is the evaporation rate. As the PDA measures volume
fluxes rather than mass fluxes, the evaporation rate was defined on a
volumetric basis: It is the difference VEVA P between the initial spray volume
flux Vo in plane x = 0 and the volume flux still passing the downstream
plane x considered, divided by the initial volume flux Vo. A global
description of the changes in the dropsize distribution due to the influence of
evaporation can be attained by considering the streamwise development of
three characteristic spray diameters: the volume undersize diameters DIO'/.
and D90 % together with the Sauter mean diameter SMD, which is a measure
of the volume-to-surface ratio of the spray. The D 90 % is the drop diameter
such that 90% of the total liquid spray volume is contained in drops of
smaller diameter. These three diameters characterize a lower, upper and
mean diameter region of the dropsize distribution and are formed by volume
flux weighted averaging of the local respective values at all measurement
points in the plane x under consideration. This gave nearly the same results
as the evaluation of these quantities from their basic definition applied to all
particles passing the plane x, as was done in the prediction.
The measured and predicted streamwise development of the integral
quantities, evaporation rate and characteristic spray diameters, is compared
in Figure 5 for the baseline case, since these are the quantities of main
interest for the combustion engineer.
It can be seen that already x = 50 mm downstream of the atomizer edge,
nearly 85% of the liquid volume is evaporated. Despite of this fast
evaporation the characteristic diameters experience but little changes over
most of the spray lifetime. Certainly, the diameter of each individual particle
in the spray will decay over the whole or a large portion of its lifetime, but
this decay is faster for smaller particles than for larger ones (see text above
Fig. 6). Depending on the relative number of large and small droplets in the
KEROSINE SPRAY EVAPORAnON 325

dropsize distribution, i.e., depending of its width and shape, this different
diameter change of the particles can lead to an increase, decrease or no
change at all of the characteristic diameters due to the course of
evaporation. In the present case little change was observed. This means
that there is a dynamic equilibrium between mass sources and mass sinks in
the diameter bins of the developing dropsize distribution over most of the
spray lifetime.
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

4.1.1. Generation and Effect of Spray Initial Conditions

Evaporation rate and development of the characteristic spray diameters may


change dramatically when altering flow parameters, in particular the initial
dropsize distribution or the spray initial temperature (Figs. 5, 7 and 16, or
see pronounced effects in Rachner et al., 1996). The knowledge of these two
quantities is the key to a successful prediction of the spray behaviour. So the
spray initial conditions used in the predictions will be discussed now:
The disintegration process of the liquid sheet begins at the atomizer edge
and might be finished about 10 millimeters downstream (as was concluded
from droplet Weber numbers based on droplet relative velocities obtained
from the predictions including spray feedback), resulting in an initial
dropsize distribution. For the computations, virtual initial conditions for
completely formed spherical droplets were assumed, starting in plane x = 0
with an approximation of the initial dropsize distribution. In the starting
plane, the given liquid volume flux was distributed in a z-range from -0.5 to
+ 0.5 mm, decreasing linearly from its maximum value at the spray
symmetry plane z = O. In prefilming airblast atomizers, the air accelerates
the droplets from the low film velocity (3 mls at slit on the atomizer for the
baseline case). So the mean initial particle streamwise velocity component
was chosen to be small (10 m/s), and mean transverse droplet velocities were
neglected. In contrast, for the droplet initial rms-velocity components, a
nonzero value (16 m/s) had to be assumed, taking into account the flapping
motion of the atomizing liquid observed by pulsed laser light sheet
visualizations. A Gaussian distribution of the initial droplet velocity was
assumed.
The fuel, which is fed at ambient temperature into the test cell, heats up
when it flows through the atomizer, which is surrounded by the hot air. For
the baseline case a fuel temperature of 453 K was measured on the atomizer
surface (Fig. 2), which was used in the predictions as the initial temperature
T PO of the droplets. The influence of T PO on the spray behaviour is addressed
in Chapter 4.2.4.
326 M. BRANDT et al.

The initial dropsize distribution for the baseline case was derived from
measurements as follows: Basically, as the initial dropsize distribution of an
evaporating spray is altered by evaporation in a manner that is not even
qualitatively known in advance, downstream measurements of the dropsize
distribution are normally little suited to extrapolate an initial distribution
for computational use. So the measurements taken closest to the atomizer
for the baseline case, which were limited by the dense spray to x = 30mm,
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

where already 77% of the initial spray volume is evaporated, could not be
used as initial dropsize distribution for the spray code. In contrast, at
ambient temperature there is nearly no evaporation of the spray and the size
distribution measured downstream of the dense spray region can be used as
a good approximation of the initial dropsize distribution. In order to find an
'equivalent' initial dropsize distribution at ambient temperature for the
baseline case, a correlation for the SMD was employed. For the small flat
prefilming airblast atomizer and the low viscosity fuel Jet A-I the SMD
produced could be correlated to the important parameters of the
atomization by (Brandt et al., 1997):

0.5
SMD ~ 10.3 . ~a.-----_ (4)
P~fr . UAir

We determined this correlation from measurements at ambient tempera-


ture, where the SMD was drawn from the volume flux weighted averaging of
five dropsize distributions measured in the plane x = 100mm. The formula
covers an atomizer loading I between 66 and 375 g/s/m, UAir between 100 and
170mis, and an air density between 2.8 and 8.4 kg/m" The exponents were
obtained by a least squares fit. The 95% confidence interval for the
exponents of I, p and UAir is 0.2 or smaller. The influence of the surface
tension of the fuel a was not investigated, so it was taken from the standard
formula of Lefebvre (1989, p. 258).
The 'equivalent' dropsize distribution at ambient temperature
(TA i r = 288 K) was measured with the same fuel loading and air density
as for the baseline case. To maintain the air density of the baseline case, an
air pressure of 3.4 bar at ambient temperature was chosen. Compared to the
surface tension of the fuel at baseline case conditions, where a fuel
temperature of 453 K near the atomizer edge was measured, the surface
tension at ambient temperature is about twice as high. To compensate for
this, the air velocity at ambient temperature was set to 170m/s compared to
120m/s for the baseline case, so that the factor aO. 5/ UAir remained constant
to get the same SMD of the size distributions under both conditions.
KEROSINE SPRAY EVAPORATION 327

Assuming that each diameter of the dropsize distribution transforms like


the SMD when changing operating conditions, the dropsize distribution
measured at the equivalent cold condition can be used as initial dropsize
distribution for the spray computation of the baseline case. This measured
size distribution is shown in Figure 6, where it is plotted as number
distribution dNjdD, volume distribution dQjdD and integrated (cumulative)
volume distribution Q(D). These curves are equivalent and each is
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

normalized by its maximum value. Volume undersize diameters can be


directly drawn from the Q(D)-curve.
The size distribution shown is based on 300000 single events, but only a
few dozens were found with droplet diameters D > 50 urn. Some of these
may be measurement errors, because of sporadic erroneous validation of
reflected instead of refracted light by the PDA electronics.
Using the measured dropsize distribution without any modifications as
the initial distribution for the spray computation resulted in the predicted
behaviour shown in Figure 7.
The evaporation rate is in good agreement with the measurement of the
baseline case, but the predicted characteristic diameters soon exhibit a
strong rise that is most pronounced and occurs at the most upstream
location for the D 90 % . This behaviour is not observed in the experiment. It is
caused by the few very large starting droplets, the diameter decrease of
which is much slower than for smaller droplets: in the baseline case the
diameter of a starting 15 urn, 50 urn and 65 urn droplet diminishes after
x = 50 mm by 11.8 urn, 4.9 urn and 4.0 urn respectively, i.e., the liquid
contained in smaller droplets evaporates faster. Thus the initially very small

r dN/dO
;f \, dO/dO
I 0(0)
0.75 I
CI>
::l
!
i;j
> i
~
i
!
0.150

...
~ !
c.es
,i
!
GO.O 80.0 '00.0

o [11m]
FIGURE 6 Unmodified, measured initial dropsize distribution for the baseline case.
328 M. BRANDT et al.

120 100

80
~

E ;g-
.2,
... 60
e.-
.l!l
Gl
E -
.>o
~
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

ltl 40 >
'6 40 w
.>
20
20

50 100 150 200


x [mm]
FIGURE 7 Baseline case, but prediction based on the unmodified initial dropsize distribution:
stream wise development of characteristic diameters and evaporation rate (symbols = measure-
ment baseline case, lines = prediction). Note the different diameter scale compared to Figure 5.

amount of liquid contained in the large droplets will begin to govern the
behaviour of the developing size distribution, if the total liquid volume has
diminished to the order of the fraction of liquid, that was initially in the
large droplets. Certainly, the evaporation rate will be governed then by the
large droplets too, but this occurs at a state, where most of the evaporation
has already happened. As finally the upper diameter end of the size
distribution governs the evaporation, the rise of the D 90 % is more
pronounced and occurs earlier than for the other two characteristic
diameters shown. So especially the development of the D90 % turnes out to
be sensitive to the shape of the upper diameter end of the size distribution.
Figure 8b shows the streamwise development of the initial dropsize
volumetric distribution dQ/dD. The plotted size distribution in each
streamwise plane was taken at the spray symmetry axis z = 0 (i.e., at
maximum of liquid volume flux density) and each distribution is normalized
by its maximum value.
It can be seen clearly how the initially very small wiggles at the upper end
of the distribution grow over the course of the evaporation process. As the
measurement did not exhibit this behaviour and the very few measured
particles above D = 50 urn (~0.01 % of 300000 events) are statistically not
significant, the upper end of the measured size distribution had to be
modified: The initial size distribution dQ/dD was approximated by a smooth
spline for diameters above 31 um, where the noise was no longer small
KEROSINE SPRAY EVAPORATION 329

compared to a smooth distribution. At 671.lm the spline was forced to reach


zero with zero gradient. This maximum diameter was derived from the
experimental observation, that in the measured size distribution for the
baseline case at x = 150 mm nearly no droplets above 551.lm were registered.
It was found then from numerical simulation that a droplet having this
diameter at x = 150mm must have had a diameter of approximately 671.lm
at the beginning. As the liquid volume contained in the splined portion of
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

the size distribution was only 1.5% and the original liquid volume above
671.lm was only 2%, the effect of this modification on the evaporation rate
was indeed minor, and the initial value of the D 90 % was changed only from
24.8 to 23.5I.lm. However, after this modification the streamwise develop-
ment of the characteristic diameters was now in quite good agreement with
the measurement (Fig. 5, compare to Fig. 7). The streamwise development
of this modified initial dropsize distribution, which was used as a basis for
all further predictions that will be shown, is given in Figure 8a. It should be
mentioned, that the initial size distribution could not be approximated
satisfactorily by a: Rosin-Rammler fit. A detailed study of the influence of
modifications on the measured initial size distribution for the baseline case is
given in Rachner (1996) and (in a more compressed form) in Brandt et al.
(1997).

4.1.2. Droplet Heat-up

Figure 9 shows the predicted heating of the droplets, which start at x = 0


with the temperature T PO = 453 K from the measurement.

FIGURE 8 Predicted streamwise development of normalized volumetric dropsize distribution


dQjdD. (a) baseline case (i.e., using the modified initial dropsize distribution); (b) baseline case,
but using the unmodified initial dropsize distribution.
330 M. BRANDT et at.

530

520 ..._._._-_._._._ ...


~

..!..o
S2'~
510
><
::J
l!! ;;::::
...~
::J 500
u
Gl

:e
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

8- 490 III
E Co
...
Gl
Gl 480
0.4
"tl
Gl
U ~
:e III
III
E
Co 470 0
0.2 c:
460

50 100 150 200 250


x [mm]
FIGURE 9 Baseline case: predicted streamwise development of droplet temperature for three
diameter classes: 6 urn ± 20%. 15urn ± 10%, 30 urn ± 5% and for the spray-mass-flux weighted
droplet temperature along with the spray mass- and volume flux.

The temperature value at a streamwise position x are number averages of


such droplets passing the plane x, which at that moment fit into one of the
three diameter classes 6 urn ± 20%, 15 urn ± 10%, 30).lm ± 5% chosen. This
gives an impression of the different heating-up times of small, medium and
large droplets in the spray. Certainly, these three curves do not represent the
heating history of a 6, 15 or 30).lm droplet, because evaporation lowers their
diameters, so that an individual droplet experiences a somewhat faster
heating than given by the three curves. In addition a 'mass weighted'
temperature curve is shown, which is the particle-mass weighted average of
the temperature of all droplets passing a plane x. This curve may globally
characterize the heating of the spray as a whole. As can be seen, most of the
spray evaporation still occurs during the heating-up of most droplets. They
finally reach a stationary end temperature of the evaporation, the wet-bulb
temperature, where the heat transported from the gas to the droplet surface
is just balanced by the energy required for evaporation. No additional
heating-up occurs after this. For the baseline case the wet-bulb temperature
turned out to be 518 K, which is 86 K below the mean boiling temperature of
604 K at a pressure of 9 bar. This 86 K -difference is a sizeable fraction of the
KEROSINE SPRAY EVAPORATION 331

initial driving temperature difference T A i r - TPO = (750-453 K = ) 297 K. It


can be concluded that simple evaporation models that consider a sequential
heating of the droplets to its boiling temperature and a succeeding
evaporation, will be considerably in error. Such models are not appropriate
for evaporation under LPP-typical conditions, and -by the way-their savings
in computer time are not substantial compared to our evaporation model.
The temperature dependency of the density of liquid hydrocarbons is
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

much stronger than e.g., for water. So the relatively small droplet heating
from 453 to 518 K lowers its liquid density already by 8%. For that reason
the normalized liquid volume flux in Figure 9 differs from the liquid mass
flux curve.

4.1.3. Effect of Spray Feedback on the Gas Phase

After atomization, the droplets experience a strong acceleration owing to


the high relative velocity compared to the air. Figure 10 shows a comparison
of measured and predicted particle streamwise velocities of the baseline case
for the 6 urn- and 30 urn diameter class.

0.8
..
~

.!..

i
0.6 R
.11
1::

.....a
0.4 .t1

tc:

50 100 150
x [mm]

. FIGURE 10 Baseline case: streamwise development of the averaged streamwise particle


velocity for two diameter classes (6 11m ± 20%, 30 11m ± 5%) along with the spray volume flu.
(symbols = measurement, solid lines = prediction in homogeneous gasfield without spray
feedback, long dashed lines = TRUST-prediction of duct flow without spray feedback, dashed
lines = TR UST-prediction of duct flow with spray feedback).
332 M. BRANDT et al.

The measured velocity values were obtained by volume flux weighted


averaging over the respective number-averaged values obtained at all
measurement points in plane x. In the predictions the velocity values in
plane x were obtained analogously as the particle temperatures above. Three
predictions are shown: first, a spray computation in a homogeneous gasfield
without spray feedback. This is the way, in which all numerical results
shown in the other figures of this paper are obtained. The other two
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

computations were performed with the TRUST-code (see Chapter 3) and


account for the developing turbulent gasfield in the duct with and without
taking into account the spray feedback on the gasfield.
In these two computations, block profiles for the flow variables were
chosen 120mm upstream of the atomizer edge as initial conditions of
the gasfield in the rectangular duct. In the test rig, this location is in the
transition piece between the round supply pipe for the hot air and the
rectangular duct. The acceleration of the spray by the air acts as a
momentum sink for the gas, which lowers the streamwise gas velocity
locally, so that predictions without feedback overpredict the droplet
acceleration. This effect can be seen to be more pronounced for the small
particles, because owing to their smaller relaxation time their driving relative
velocity (UAir - up) is already much smaller at x-planes where the lowering of
the gas velocity due to spray feedback mainly occurs. So they react more
sensitively to the changes in the gas velocity there than larger particles.
For the baseline case it can be seen from Figure 10, that even the prediction
with spray feed-back still overpredicts the particle velocities somewhat. The
reason might be, that the disturbance (wake) of the flow field by the atomizer
body was not taken into account. From a comparison of the TRUST-
computations with and without spray feedback, a maximum local lowering of
the gas streamwise velocity due to spray acceleration of 10 m/s at the duct
center at x = 9 mm was found. The maximum local lowering of the gas
temperature due to spray heating and evaporation was 47 K at x = 27 mm.
Figure 10 reveals that the differences in the droplet acceleration,
computed with different assumptions about the gasfield, had very little
effect on the evaporation rate. Also the computed streamwise development
of the characteristic diameters (not shown) is nearly identical for the two
TRUST-computations and differs not much from the characteristic
diameters from the computation in the homogeneous gasfield. So neglecting
the spray feedback is a good approximation in our case, except if analysis of
droplet dynamics is of main interest. These results bear the justification of
performing most of the numerical spray analysis in a homogeneous gasfield
without spray feedback, resulting in substantial savings of computer time.
KEROSINE SPRAY EVAPORATION 333

4.1.4. Spray Dispersion and Fuel Vapour Mixing


The measured spray dispersion and fuel vapour mixing perpendicular to the
atomizer plane (z = 0) is shown in Figure II, where the streamwise
development of the halves of the profile widths that contain 10%
respectively 90% of the measured fuel flux (zo.), ZO.9) together with the
standard deviation Zrms are presented. To compare the local liquid volume
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

fluxes measured with the PDA with the relative vapour concentrations based
on the light extinction technique that integrates along the laser beam, the
liquid volume flux measurements were integrated correspondingly.
As can be seen, the transverse spray penetration was small compared to
the streamwise position: the standard deviation Zrms at x = 150 mm was only
4 mm. Moreover, a slightly larger penetration of the fuel vapour compared
to the liquid spray was observed. As 80% of the vapour in the baseline case
is set free until x = 40 mm, the transverse spreading of the vapour is
governed in that range by the dispersion of the massively evaporating spray.
This situation changes farther downstream, where the remaining portion of
the liquid spray acts only as a weak source of vapour, so that their
transverse spreading is decoupled then, and any differences between the
transverse spreading of spray and fuel vapour are caused by particle inertia
effects. These are not negligible for the present. case, which is demonstrated
by the fact that only with proper values for the initial particle rms-velocities

E 10
.ss:
-
'C
~
8 -

6
--
.
c:
0
::J
4

-
.c
.;:
Ul
is
2
0
0 20 40 60 80 100
--------
120 140 160
x [mm]

Spray Vapour
ZO.1 Zrms ZO.9 ZO.1 Zrms ZO.9
--&- ~ --B- -0- -+- -.
Prefilmer 8 mm, P=9 bar,TAir=750 K, 1=125 g/s/m

FIGURE II Measurement baseline case: streamwise development of different measures


characterizing the half width of the transverse spray- and vapour flux distributions.
334 M. BRANDT et al.

a good agreement between measured and computed liquid spray dispersion


was obtained (Rachner et al., 1996). Therefore it can be concluded, that the
measured small differences between the transverse penetration width of
spray and vapour is due to compensating effects of particle inertia and will
change, if the turbulence properties of the gasfield (turbulent time- and
length scales) change.
Accordingly the experiments showed that the parameter variations of the
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

baseline case, which had very little effect on the gasfield turbulence, had a
very small influence on the droplet dispersion and fuel vapour mixing. Spray
predictions of the baseline case with and without using a turbulent
dispersion model demonstrated that turbulent gas velocity fluctuations
had but a very small influence on the evaporation rate: dispersion and
evaporation turned out to be nearly uncoupled. This is consequence of the
small gradients of the gasfield in the duct: despite of different individual
droplet paths caused by the turbulent dispersion, the droplets experience
about the same gasfield surroundings. However gas turbulence has an
influence on the particle dispersion, therefore a turbulent particle dispersion
model is necessary to predict the transverse spreading of the droplet cloud.
It can be summarized that due to the small gradients in the time-averaged
gasfield, the transverse spreading of spray and vapour was only small. The
turbulent features of the gasfield were important for the spray dispersion
and vapour mixing, but had very little effect on the evaporation rate. A
doubling of the penetration width was achieved by introducing turbulence
generators in the duct (not shown here, but in Brandt et al., 1997).

4.2. Influence of Operating Parameters on the Baseline Case

In this chapter the influence of air pressure, air temperature, fuel initial
temperature, air velocity and atomizer loading is studied. These are the
operating parameters of the combustor which change during a flight cycle of
an aeroengine, and are subject to design choices of the combustion engineer.
Moreover the question whether a tractable, single-component evaporation
model can describe the behaviour of a multicomponent fuel like kerosine is
addressed, and the sensitivity of the model to (supplier dependent) changes
in fuel volatility is investigated (Chapter 4.2.3).

4.2.1. Influence of Air Pressure

The measured influence of air pressure on the atomization and evaporation


of kerosine Jet A-I is shown in Figure 12. At an air temperature of 750 K, an
KEROSINE SPRAY EVAPORA nON 335

~ 90'
,.g
i!.....
.>080'
- II.
;;
70·

.>w 60
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

50 150
0 30 60 90 120

25

20
E 15 .
2:
c
:;: 10
lfl

a
a 30 60 90 120 150
x [mm)

P =3 bar P =6 bar P =9 bar P =12 bar P =14.5 bar


---- -+- -+- ~
Prefnmer amm, TAir=750 K, u=120 m1s,I=l25gJslm

FIGURE 12 Measured influence of pressure on streamwise development of Sauter-diameter


and evaporation rate.

air velocity of 120m/s and a fuel loading of 125g/s/m (i.e., baseline case
conditions) the air pressure in the test cell was varied from 3 bar to 14.5 bar.
As can be seen, raising the air pressure leads to a distinct increase of the
evaporation rate. Whereas at 3 bar at x = 150mm downstream of the
atomizer edge only 73% of the fuel volume was evaporated, at 14.5 bar
already at x = 60 mm evaporation rates of 97% we're found. It should be
noted here that the liquid volume flux measurements have a margin of the
relative error of + 10% and - 14%, and therefore the measured evaporation
rate (= (Vo - V)/ Vol of 73% is between 70 and 77%, whereas at 97% it lies
between 96.7 and 97.4%. This increase of the evaporation rate with the air
pressure is caused to a large extend by the better atomization at higher air
density, which is proportional to the pressure at constant temperature.
Measurements at ambient temperature, which led to the correlation of Eq.
(4), showed a decrease of the initial SMD with the air density by the factor
PA?~3. Indeed Figure 12 shows that the SMD of the spray decreases in
336 M. BRANDT et at.

the first two measurement planes with the air pressure. For a constant
pressure the SMD remains almost constant during evaporation. Only at the
highest pressures at the last measurement planes an increase in SMD is
observed. This might be caused by the very few large droplets in the size
distribution, which begin to dominate the characteristic diameters (cf
Chapter 4.1.1), as more than 99% of the spray is already evaporated.
Other factors influencing the initial spray conditions and enhancing the
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

evaporation are caused by the kerosine temperature. The fuel is fed into the
test cell at ambient temperature, but it is heated up when it flows through
the atomizer surrounded by the hot air. At higher air densities the heat
transfer to the atomizer housing and the liquid is increased. The experiments
showed that for the operating conditions presented in Figure 12 the fuel
temperature near the atomizer edge increases from 400 K to 480 K, if the air
pressure is raised from 3 bar to 14.5 bar. Such an elevated spray initial
temperature augments the succeeding evaporation (see Chapter 4.2.4) and in
addition lowers the surface tension. The surface tension of kerosine Jet A-I
exhibits an almost linear decrease with temperature: It diminishes from
0.015 Njm to 0.010 Njm when the fuel temperature increases from 400 K to
480 K (Rachner, 1998) and this in turn reduces the initial dropsize diameter
(see Eq. (4», which enhances evaporation.
The influences of the increase of the air pressure from 3 bars to 14.5 bars
on the initial spray conditions may be quantified as follows:

• a decrease to 62% of the initial SMD due to the influence of the air
density on atomization (Eq. (4»
• an increase of the initial fuel temperature from 400 K to 480 K
• a decrease to 82% of the initial SM D due to the influence of the fuel
temperature on the surface tension at atomization (Eq. (4»
A separate study of the influence of the air pressure on evaporation itself
could not be conducted experimentally, because the air pressure was coupled
with fuel heating and atomization. However the effect on evaporation itself
was separated by the numerical spray simulation. Figure 13 shows three
computations made with the same initial dropsize distribution as used for
the baseline case, but at varying air pressures.
From this figure it is found in our case that pressure deteriorates tbe
evaporation process, but only moderately: at x = 30 mm the evaporation.
rate diminishes from 83% at 3 bar to 72% at 14.5 bar. Further
downstream at x = 100mm in all cases evaporation rates of more than
98% are found. This influence of raising the pressure is a result of several
interacting effects:
KEROSINE SPRA Y EVAPORAnON 337

100 _-------------,50
,~.~.~.-r.:J' - - p=3 bar
90
I," ----- p= 9 bar
80 I~"" ..........".. p =14.5 bar 40
,..-"'..-
70 ': ...............................................................
,-'" ....-
1;'
~
---·-D~~~---- E
~ 60
o
,:
....s
.. - 30 ~

•> ¥.:.;....:,.; .:,:::..::.;:......


Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

.... 50 ..

~
w
.>

50 100 150 200 250


x [mm]
FIGURE 13 Predicted influence of pressure only on evaporation process itself: streamwise
development of characteristic diameters and evaporation rate (all other parameters of baseline
case left unchanged).

As the raising of pressure increases the air density, which in turn increases
the particle Reynolds number and therefore lowers its Stokes number, the
particles are faster accelerated and reach a given streamwise position in a
shorter time, so their residence time and time to evaporate is reduced.
Moreover, the elevated particle Reynolds number enhances heat and mass
transfer coefficients at the droplet. Also the fuel has to be heated up to a
higher wet-bulb temperature of 542 K at 14.5 bar compared to 469 K at 3
bar and the heat-up period of each droplet takes a longer portion of its
lifetime. Finally, the temperature dependence of fuel properties (e.g., liquid
fuel density and heat of evaporation decrease significantly with temperature)
is of influence, when droplet heating and wet-bulb temperature change.
So it can be concluded, that the measured increase of the evaporation
rates with the pressure is mainly caused by the finer spray accompanied by a
higher spray initial temperature. This effect is somewhat weakened by the
slower evaporation mechanism at higher pressures.

4.2.2. Influence of Air Temperature


At an air pressure of 9 bar, an air velocity of 120m/s and a fuel loading of
125g/s/m the air temperature was increased from 650 K to 850 K. This
338 M. BRANDT et al.

increase of the air temperature has a distinct effect on evaporation. As the


measurements (symbols in Fig. 14) reveal, for TA i , = 850 K at x = 60mm
more than 99% of the initial spray volume was evaporated, whereas at
650 Keven 150mm downstream of the atomizer the evaporation rate was
only 91%.
According to Eq. (4) there are two counteracting effects of the air
temperature on the atomization: a higher air temperature lowers air density
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

and this alone would give a coarser spray. In contrast, the higher air
temperature induces a higher spray initial temperature by enhanced preheat
of the fuel from the air through the atomizer housing. This higher spray
initial temperature enhances the succeeding evaporation but also reduces the
surface tension, which gives a finer spray. The surface tension effect turned
out to be somewhat more pronounced than the influence of the air density,
so that the 8MD of the spray produced by the atomizer exhibits a slight
decrease with air temperature.
These different effects on the initial spray conditions can be quantified as
follows, when raising air temperature from 650 to 850 K:
• an increase of the initial 8MD to 108% caused by the effect of the reduced
air density on atomization (Eq. (4»

100

80

;,!!
e:. 60
0

'c;.
"
>
w
40
.>
20.'"i-'-' 10
,I
I
I

50 100 150 200


X [mm]
FIGURE 14 Influence of air temperature on streamwise development of Sauter-diameter and
evaporation rate (symbols = measurement, lines = predictions varying the baseline case:
dashed lines = TAi,=650 K, long dashed lines = T A i,=650 K and T PO, DPO from measurement,
solid lines = TAir=850 K, dash-dotted lines = TA ir = 850 K and T PO' D PO from measurement).
KEROSINE SPRAY EVAPORATION 339

• an increase of the initial fuel temperature from 414 K to 480 K, whereby


the dependence between initial fuel temperature and air temperature was
about linear
• a decrease of the initial SMD to 84% caused by the effect of the fuel
temperature on the atomization via surface tension (Eq. (4))
The influence of the air temperature on evaporation itself could not be
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

separated by the experiment. So numerical simulations at 650 K and 850 K


were accomplished and some results are presented in Figure 14 together with
the measurements.
Figure 14 shows two sets of variations of the baseline case: First, only the
value of the air temperature was changed to 650 and 850 K (same spray
initial conditions as baseline case!). Second, the changes in the spray initial
conditions were approximated additionally: The respective measured initial
spray temperature (414 K at T A ir = 650 K, 480 K at TA;r = 850 K) was used
and the dropsize distribution of the baseline case was scaled by multiplying
the diameter of all starting computational droplets by a respective factor
(1.06 at TA;r = 650 K, 0.96 at T A ir = 850 K). This factor is the ratio of the
SMDs at the respective condition compared to the baseline case according
to Eq. (4), and takes into account the change of the air density as well as the
surface tension. By means of these two sets of computations it is possible to
distinguish between the influence of air temperature on the droplet
evaporation process itself and the additional effect caused by the changed
spray initial conditions. It is found that the former effect on the evaporation
rate is more pronounced than the latter. The computation with approxi-
mated spray initial conditions is in good agreement with the measurements
and reveals an already significant development (decrease) of the SMDs from
x = 0 to x = 30mm.

4.2.3. Sensitivity to Fuel Volatility

In order to estimate the effect of the scattering in volatility among Jet A-I
charges, we changed the underlying vapour pressure relation for the
prediction of the T A ir = 650 K-case: The original vapour pressure curve
yields a mean boiling temperature of 478 K at I bar for Jet A-I (note that in
the TA;r = 650 K-case the droplets were heated from 414 to 503 K, so the
vapour pressure relation was applied in this range near the normal boiling
temperature). At this normal boiling temperature typically 48% of the
Jet A-I liquid volume is recovered in the ASTM-distillation apparatus
(Rachner, 1998). Now the vapour pressure curve was modified such that
340 M. BRANDT et al.

the boiling temperature at I bar was 19 K higher, which is just the maximum
deviation found among the many individual distillation curve data for
Jet A-I compiled in Rachner (1998) at the 50% -recovered temperature
compared to the typical value for Jet A-I. This temperature of 497 K
corresponds to the 76% -recovered temperature of the typical distillation
curve of Jet A-I. With this modified vapour pressure relation (and the initial
dropsize distribution of the baseline case scaled with the SM D-ratio of 1.06),
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

a rise in the wet-bulb temperature by 15 K to 518 K and an evaporation rate


of 84.1% at x = 100mm was observed. Compared to a value of 88.5%
obtained with the original vapour pressure relation, this is only a small
influence exerted from the possible scattering in volatility on the
evaporation rate.
Jet A-I is a multicomponent fuel consisting mainly of hydrocarbon
compounds with 9 to 17carbon atoms (Rachner, 1998). At fast evaporation,
the local depletion (arising from the different evaporation behaviour of the
species) of volatile components at the surface of the multicomponent droplet
is hindered by the diffusion resistance in the droplet interior. Jn the limit an
'onion peel'-type evaporation occurs, where the multicomponent fuel
behaves like a single-component fuel. In contrast at slow evaporation
(e.g., at low air temperatures) there is more time for supplying the droplet
surface from inside with volatile components, and the selective character of
the multicomponent fuel evaporation is more pronounced. The good
agreement between our computations and the experiment under LPP-typical
conditions at T A i r = 850,750 and 650 K indicates that the single-component
treatment of the Jet A-I evaporation is justified in that range. This is
confirmed by the computational study of Kneer (1993) of the evaporation of
a two-component heptane-dodecane droplet compared to a single-
component n-decane droplet, where only a small influence of the selective
evaporation on the evaporation rate at T A i r = 800 K was stated.

4.2.4. Influence of Spray Initial Temperature

In order to isolate the effect of the initial spray temperature on the droplet
evaporation process itself, two predictions of the baseline case were
performed, in which only the initial spray temperature T PO was varied
from the measured value 453 K. All other parameters remained unchanged.
Two extreme values were chosen as the spray initial temperature: the fuel
supply temperature (288 K) into the test rig, and the wet-bulb temperature
(518 K). The latter value implies, that the droplets experience no heating at
all during evaporation.
KEROSINE SPRAY EVAPORATION 341

Figure 15 reveals the strong influence of the spray initial temperature on


the evaporation. At x = 70 mm the evaporation rate of the 288 K -case
reaches 77.2%, whereas for 518 K the respective values is already 99%,
revealing a strong effect of droplet heating on evaporation. The negative
values of the evaporation rate for the 288 K -case computed in the range
x < 14mm are caused by the volume expansion of the droplets due to the
lowering of their liquid density, which is at that state not yet compensated
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

by the (still slow) evaporation. However the underlying 'uniform


temperature model' used in the spray code to describe droplet hearing
tends to underpredict the droplet surface temperature and therefore a slower
onset of evaporation.
In a premix duct with an airblast atomizer the predicted strong increase of
evaporation with the spray initial temperature will be further enhanced by
the lowering of the surface tension, which leads to a finer spray. So, if a
significant part of fuel heating could be achieved in the fuel path, the
prevaporizer length could be reduced considerably. However the permissible
extent of preheating is limited by the thermal stability of the fuel (see
Rachner, 1998).

4.2.5. Influence of Air Velocity


According to Eq. (4) the air velocity is the most important parameter for
airblast atomization. But it is also important for the succeeding droplet

100 l--:::~~tDi~--=:::===~------,
40

80
30
~

E
~
20 0
::iE
.. - en
10
o

100 150
x [mm]
FIGURE 15 Predicted influence of droplet initial temperature on streamwise development of
Sauter-diameter and evaporation rate (all other parameters of baseline case left unchanged).
342 M. BRANDT et al.

evaporation process, because a higher relative velocity between gas and


droplet raises its particle Reynolds number. This in turn enhances the heat
and mass transfer at the droplet and in addition influences its acceleration
and terminal velocity and therefore its residence time. These relative velocity
effects considerably influence the observed rapid evaporation in the baseline
case near the atomizer edge. Figure 16 shows the prediction for the baseline
case together with two variations: first, the air velocity was reduced from
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

120 mls to 100m/s. Second, the effect on the initial dropsize distribution was
approximated additionally by multiplying the diameter of all starting
computational droplets by a factor of 1.2, which is the ratio of the SMDs at
UAir = lOOmis and 120m/s according to Eq. (4). This step implies that all
initial characteristic diameters increase by this same factor.
It can be seen that the isolated effect of lowering the air velocity on the
evaporation process resulted in faster evaporation. Obviously the deteriora-
tion of heat and mass transfer at the droplets for UAir = 100mls was
overcompensated by the longer residence times. However, when additionally
taking into account the increased initial dropsizes, the evaporation becomes
slower than in the baseline case. This confirms the importance of the initial
dropsize distribution on evaporation.

100 .,;..;;_--------------,50

90

80 .......................................................... 40

70 ..... 0. 0%
...... ...... .. _---.
~
:,!! 60
e:- .>:
~

E
30 .;;!,
...
-
o
"> 50
.....~.fJI.P. · ~E
!1l
>
w 40 ...................... 20 .S!
~:.;.:.;:.;.:.;:.:=c:.=c=-"1 "C
">
30

20

10

0
0 50 100 150 200 2S0
X [mm]

FIGURE 16 Predicted influence of air velocity on streamwise development of characteristic


diameters and evaporation rate (full lines = baseline case (i.e., UAi, = 120m!s), dashed lines =
baseline case, but UAir = 100mis, dotted lines = baseline case, but UAir = 100mls and
DPO = 1.2D PO baseline)'
KEROSINE SPRAY EVAPORATION 343

For lower air velocities it should be remembered here that the SMD-
correlation given in this work was validated only for air velocities between
100 and 170 tn]«. Moreover it should be emphasized that the simple
transformation of the initial dropsize distribution by a factor derived from a
SM D-correlation presumes that the basic shape of the size distribution does
not change. However, other mechanisms for liquid disintegration and
droplet breakup, as reviewed in Lefebvre (1989, Chapter 2), become
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

increasingly important at lower air velocities, which might increase the


upper range of the dropsize distribution drastically, and in turn would have
a large effect on evaporation.

4.2.6. Influence of Fuel Loading of the Atomizer

The influence of the fuel loading of the atomizer on atomization and


evaporation was studied experimentally with the larger atomizer (Fig. 2), the
prefilming plane of which had to be mounted vertically in the duct.
Therefore the light extinction technique (see Chapter 2.3) could not be
applied to get fuel vapour profiles perpendicular to the prefilmer plane.
However due to the larger prefilmer width (25 mm rather than 8 mm) the
effects of reduced particle densities at the lateral borders of the spray are
negligible, so that this atomizer represents a better two-dimensional
approximation and therefore was the more suitable atomizer to study the
effect of an increased fuel flow rate. The experimental results are shown in
Figure 17.
An increased fuel flow rate leads to an increase of the initial SMD of the
spray and a decreased evaporation rate. A comparison with Figure 12,
where results with the small atomizer are shown, reveales that in this case
the slit height (0.3 mm for the larger atomizer rather than 0.1 mm) and the
atomizer widths do not playa dominant role on atomization, because at the
same pressure and about the same fuel loading ratio almost the same
dropsizes and evaporation rates were measured. I
The effects of an increased fuel flow rate from 83.3 to 333 glslm (Fig. 17)
on the initial spray conditions may be summarized as follows:
• an increase of the initial SMD to 152% caused by the increased fuel
loading (Eq. (4»

1 For the investigated operating conditions the highest velocity of the fuel leaving the
atomizer slit was 3m/s, so that the relative velocity between the fuel and the air -an important
factor influencing airblast atomization- was nearly unaffected by the atomizer design and the
fuel flow rate.
344 M. BRANDT et al.

I--~:=::====~:::::=:;::::====::::;;;.l
7
~
o
100
90· ...
-
">
Q.
~ 70'
80

.>
60 .
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

50
0 30 60 90 120 150

25

20 -
E
.:!: 15 -
c
:; 10 .
1Il

0
0 30 60 90 120 150
x [rnm]

1= 83.3 glslm 1=167glslm 1=250 glslm 1=333 glslm


~ ........ --+- -----
Prelilmer 30 mm, p= 9 bar, T 1:1 750 K. u= 120 mls

FIGURE 17 Measured influence of atomizer fuel loading on streamwise development of


Sauter-diameter and evaporation rate (larger atomizer used).

o a decrease of the initial fuel temperature from 480 K to 400 K (measured


for the small atomizer, but might be similar for the larger atomizer)
• an increase of the initial SMD to 122% caused by the increased surface.
tension of the fuel (Eq. (4))
These effects tend to decrease the evaporation rates with increased fuel
loading. On the other hand the evaporation rate is somewhat enhanced by
the higher residence times of the fuel droplets at higher fuel loading. This is
caused by the additional momentum loss of the air that had to accelerate a
larger amount of fuel. Compared to the mean streamwise particle velocities
for the baseline case (Fig. 10, fuel loading 125 gjmjs), a typical reduction of
about 10 mjs for the 6 urn diameter class and about 6 mjs for the 30 urn class
was measured in all four measurement planes for the larger atomizer at a
fuel loading of 333 g/s/m.
KEROSINE SPRAY EVAPORATION 345

The comparison of calculations with and without feedback of the spray


on the gas field (Fig. 10) showed, that the spray feedback, which leads to a
reduction of the particle velocities, had only minor influence on the
streamwise development of dropsize distribution and evaporation rate.
Therefore it can be concluded that for the increase of the fuel loading in
Figure 17, the decrease of the evaporation rate is mainly caused by the
coarser spray produced and the lower spray initial temperature rather than
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

by the risen spray feedback on the gasfield.

5. CONCLUDING REMARKS

Evaporation and mixing of kerosine emerging from a flat prefilming airblast


atomizer was studied experimentally and numerically in an optically
accessible, rectangular duct at conditions relevant for lean premixed and
prevaporized combustion. Liquid phase properties were measured by Phase-
Doppler anemometry and fuel vapour profiles were determined by an
infrared light extinction technique. Computations based on the Lagrangian
particle tracking technique with and without taking into account the
feedback of the spray on the gas phase. The evaporation model
approximates the Jet A-I as single-component fuel and presumes a uniform
temperature distribution in the droplet.
It was found that at pressures above 9 bar, air temperatures above 750 K
and a bulk air velocity of 120m/s the kerosine spray evaporates nearly
completely within 100mm distance from the atomizer edge. Within the
observable pathlength of 150mm, no autoignition occurred. Spray
dispersion and fuel/air mixing were not much affected by the variations of
the operating conditions, but were small. Although a doubling of the
transverse penetration width was achieved by introducing turbulence
generators in the duct, more effective means to enhance mixing in LPP-
premixers have to be used. Most promising to us is swirling air or a cross
injection of the fuel. Respective work is in progress in our laboratories.
Good agreement was found between the numerical predictions and
experimental results, after the fuel temperature on the prefilmer had been
measured and the initial dropsize distribution had been obtained with the
aid of measurements at ambient temperature and a SMD-correlation that
was fitted to the atomizer used.
The comparison of numerical and experimental results of parametric
variations of the operating conditions showed strong effects on the
evaporation history of the spray. Moreover, the relative importance of the
346 M. BRANDT et at.

direct influence of the parameters on the droplet evaporation process itself


was separated from their influence on the atomization, which then in turn
affects the evaporation.
The strong effect of operating parameters (pressure, mass flow, air and
fuel temperature) on atomization and/or evaporation means that evapora-
tion length required in LPP-premixers is sensitive to given operating
conditions of the gasturbine. Beside the air temperature, which had the
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

largest effect on evaporation rates, the initial spray temperature had a


significant influence. So a controlled preheat in the fuel path can enhance
evaporation considerably. Airblast atomizers produce a much finer spray
with increased air velocity and pressure. Both had an indirect but strong
effect on evaporation through atomization quality, whereas their effect on
the evaporation process itself was of minor importance in our case. Spray
feedback on the gasfield had no dominant effect on evaporation. The
experimentally found decrease of the evaporation rates with an increased
fuel flow rate is mainly caused by its influence on the spray initial conditions.
For an accurate numerical simulation of particle dynamics the feedback
of the spray on the gas field had to be considered, but the evaporation
history and droplet dispersion could be described well without taking
account the feedback of the spray on the gas phase. The standard single-
component Lagrange spray model-without resolving the droplet interior
numerically-could be applied with good results even to the multi-
component fuel Jet A-I under conditions relevant for gasturbine combus-
tors. However it turned out that proper estimations or measurements of the
initial spray conditions, especially the dropsize distribution and the spray
initial temperature were important for the accuracy of the spray simulation.

Acknowledgements

The present study was financially supported by the European Community


under CEC Contract n": BRPR-CT95-0122, which is gratefully acknow-
ledged. Also valuable discussions with Dr. C. Hassa at DLR are appreciated.

NOMENCLATURE

Symbol Unit Meaning

ap Im/s 2] droplet acceleration


bl [urn] burst length
Bo H Bond number of droplet: Bo = ppapD~/a
CD H drag coefficient of the droplet
KEROSINE SPRAY EVAPORAnON 347

ds [11 m] fringe spacing


D,D p [11 m] droplet diameter
De - z [11m] Gaussian beam diameter
dN/dD H normalized dropsize number distribution
dQ/dD [ -] normalized dropsize volume distribution
D 1o% [11 m] volume undersize lO%-diameter of the spray
D 90 % [11 m] volume undersize 90%-diameter of the spray
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

I [g/s/m] fuel loading (per length of atomizer edge)


L" L z [11 m] lengths in measurement volume (see Fig. 4)
L turb [mm] turbulent macro length scale
p [pa] pressure in the duct
Q(D) [ -] normalized, integrated dropsize volume distribution
s [11m] slit width
SMD [11 m] Sauter diameter (= volume/surface ratio) of the spray
r.: [K] air inlet temperature
TPO [K] initial droplet temperature at x = 0
U, v, w [m/s] velocity components in x-, yo, z- direction
UAir [rn/s] bulk air velocity in the duct
3/s]
Vex) [m volume flux of the spray in plane x
3/s]
Vo [m volume flux of the spray in plane x = 0
3/s]
VEVAP [m evaporated spray volume flux (:= Vo - Vex))
VEVAP/ Vo [%] evaporation rate
x,y, z [mm] cartesian coordinate system in the premix duct
(its origin is in the middle of the atomizer edge, x is
streamwise direction of the duct, x = 0 is streamwise
position of atomizer edge, z = 0 is plane of small
prefilmer, duct sidewalls are at z = ± 20 mm and
y = ± 12.5mm)
Zrms [mm] standard deviation of liquid or vapour flux profil in
z-direction
ZO.I [mm] 10%-width of half liquid or vapour flux profil in
z-direction
ZO.9 [mm] 90%-width of half liquid or vapour flux profil in
z-direction

Greek symbol Unit Meaning

'P [0] sea ttering angle


A [11 m] wavelength
p [kg/rrr'] density
[N/m] surface tension of the liquid fuel
348 M. BRANDT et al.

Subscript Meaning

Air air in the duct


EVAP evaporated
P droplet
rms standard deviation (rms value)
o initial state at x = 0
Downloaded by [University of Tennessee, Knoxville] at 04:27 12 May 2013

References

Bliirncke, E., Brandt, M., Eickhoff, H. and Hassa, C. (1993) Particle dispersion in highly
swirling, turbulent flows. Particle and Particle Systems Characterization, 10, 182-190.
Brandl, M., Hassa, C., Kallergis, K. and Eickhoff, H. (1994) An experimental study of fuel
injectors for premixing ducts. lCLASS-94, Rouen, France, June 1994.
Brandt, M. (1995) Liquid and gaseous fuel measurements in a premix duct. PARTEC 95, l lth
European Conferenee of /LASS Europe, Niirnberg, ISBN 3-921590-34-5.
Brandt, M., Hassa, C. and Rachner, M. (1997) Messung der Zwei-Phasen-Strornung
(Tropfengeschwindigkeit und -grobe) in einer Olvormischstrecke. Final Report TURBO-
FLAM Phase II-Vorhaben 3.2.t.8, Institute for Propulsion Technology, DLR, Cologne,
April 1997.
Clift, R., Grace, J. R. and Weber, M. E. (1978) Bubbles, drops and particles. Academic Press,
New York, ISBN 0-12-176950-X.
DANTEC (1992) PDA User's Manual. DANTEC Measurement Technology, Tonsbaken 18,
Dk-2740, Skovlunde, Denmark.
Drallmeier, J. A. (1994) Hydrocarbon vapor measurements in fuel sprays: a simplification of the
infrared extinction technique. Applied Optics, 33(30), 7175-7179.
Edwards, C. F. and Marx, K. D. (1992) Analysis of the ideal Phase-Doppler system: limitations
imposed by the single-particle constraint. Atomization and Sprays, 2, 319-366.
Fitzky, G. and Bohn, D. (1995) Messung der komplexen Brechungsindizes von Heizol EL und
Kerosin in Abhiingigkeit von Druck- und Temperatur. Bericht, Lehrstuhl und lnstitut fur
Dampf- und Gasturbinen, RWTH Aachen, Oct. 1995.
Jansen, M., Schulenberg, T. and Waldinger, D. (1991) Shop test result of V64.3 gas turbine.
ASME 91-GT-224, Orlando, Florida, June 3-6, 1991.
Kneer, R. (1993) Grundlegende Untersuchungen zur Spriihstrahlausbreitung in hochbelasteten
Brennraurnen: Tropfenverdunstung und Spriihstrahlcharakterisierung. Dissertation, In-
stitut flir Thermische Stromungsrnaschinen, Universitat Karlsruhe.
Lefebvre, A. H. (1989) Atomization and sprays. Hemisphere Publ. Corp., New York, ISBN 0-
89116-603-3.
Rachner, M. (1996) Zum EinfluB der Anfangstropfengrobenverteilung auf die Sprayausbreitung
im LPP-Vormischkanal. Internal Report DLR-IB 325-14-96, Institute for Propulsion
Technology, DLR, Cologne, Sept. 1996.
Rachner, M., Brandt, M., Eickhoff, H., Hassa, c., Braumer, A., Kramer, H., Ridder, M. and
Sick, V. (1996) A numerical and experimental study of fuel evaporation and mixing for lean
premixed combustion at high pressure. 26th Symp. (Int.) on Combustion, held at Naples,
Italy, July 28 - August 2 1996, The Combustion Institute, pp. 2741-2748.
Rachner, M. (1998) Die Stoffeigenschaften von Kerosin Jet A-I. DLR-Mitteilungen 98-01,
DLR, Cologne, March 1998.
Schutz, H., Eickhoff, H., Theisen, P., Griebel, P. and Koopman, J. (1997) Analysis of the
mixing zone of an air staged combustor. 13th Int. Sympos. on Airbreathing Engines.
(lSABE), Chattanooga, Tennessee, Sept. 7-12, 1997, International Society For Air
Breathing Engines.
Shampine, L. F. and Gordon, M. K. (1975) Computer solution of ordinary differential
equations- The initial value problem. W. H. Freeman and Company, San Francisco,
Spadaccini, L. J. and TeVelde, J. A. (1982) Autoignition characteristics of aircraft-type fuels.
Combustion and Flame; 46, 283-300, The Combustion Institute.

You might also like