Experimental and numerical investigation of the primary breakup of an airblasted liquid sheet

Experimental and numerical investigation of the primary breakup of an airblasted liquid sheet

Accepted Manuscript Experimental and numerical investigation of the primary breakup of an airblasted liquid sheet K. Warncke, S. Gepperth, B. Sauer, ...

10MB Sizes 0 Downloads 40 Views

Accepted Manuscript

Experimental and numerical investigation of the primary breakup of an airblasted liquid sheet K. Warncke, S. Gepperth, B. Sauer, A. Sadiki, J. Janicka, R. Koch, H.-J. Bauer PII: DOI: Reference:

S0301-9322(15)30214-7 10.1016/j.ijmultiphaseflow.2016.12.010 IJMF 2536

To appear in:

International Journal of Multiphase Flow

Received date: Revised date: Accepted date:

19 December 2015 1 December 2016 10 December 2016

Please cite this article as: K. Warncke, S. Gepperth, B. Sauer, A. Sadiki, J. Janicka, R. Koch, H.-J. Bauer, Experimental and numerical investigation of the primary breakup of an airblasted liquid sheet, International Journal of Multiphase Flow (2017), doi: 10.1016/j.ijmultiphaseflow.2016.12.010

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Highlights • Identical breakup phenomena are observed in experiment and simulation

CR IP T

• Ligament and bag formation in the vicinity of the atomizing edge

• Efficient characterization of the liquid sheet by tracking the phase interface • 3D-DNS data is analyzed as seen from the top corresponding to shadow images

• Droplet size distributions provided by eDNS are in good agreement with

AC

CE

PT

ED

M

AN US

the experiment

1

ACCEPTED MANUSCRIPT

Experimental and numerical investigation of the primary breakup of an airblasted liquid sheet

CR IP T

K. Warnckea,∗, S. Gepperthb , B. Sauera , A. Sadikia , J. Janickaa , R. Kochb , H.-J. Bauerb a Department

of Energy and Power Plant Technology, Technische Universität Darmstadt, 64287 Darmstadt, Jovanka-Bontschits-Straße 2, Germany b Institut für Thermische Strömungsmaschinen, Karlsruher Institut für Technologie, 76135 Karlsruhe, Kaiserstraße 12, Germany

AN US

Abstract

The primary breakup of airblast atomization is governed by complex mechanisms and is still not well understood. In recent years high speed shadowgraphy experiments and Direct Numerical Simulations of prefilming airblast atomization have been performed independently. In this paper detailed results of a

M

combined experimental and numerical study are presented. A single operating point of a planar prefilming airblast atomizer is investigated, based on a spatial resolution of 10 µm and a consistent analysis of the liquid film in both the exper-

ED

imental and the numerical studies. For the analysis the three-dimensional DNS data is projected on a plane, corresponding to the data obtained by shadowgraphy. The experiment is characterized by back light illumination in conjunction

PT

with particle and ligament tracking velocimetry. A Depth of Field correction is applied to further improve the measurement accuracy. For the numerical inves-

CE

tigation the embedded DNS approach is utilized: The primary breakup region is simulated with a highly resolved DNS, embedded in a coarser Large Eddy Simulation. The comparison comprises a phenomenological discussion of the

AC

disintegration process and quantitative results. Distributions for the breakup length, the liquid film deformation velocity, the droplet sizes and velocities are presented. The results are in good agreement and confirm the applicability of ∗ Corresponding

author. Tel. +49 6151 1628906; fax +49 6151 28900 Email address: [email protected] (K. Warncke)

Preprint submitted to International Journal of Multiphase Flow

February 3, 2017

ACCEPTED MANUSCRIPT

the embedded DNS and the particle and ligament tracking velocimetry for the analysis of the primary breakup of airblast atomization. Shadowgraphy, embedded DNS, aircraft engine, primary

atomization, ligament tracking, droplet tracking 1. Introduction

CR IP T

Keywords:

The air traffic has doubled every 15 years since the 1970s and is expected to continue this growth for at least the next 15 years (Airbus (2013); Boeing

(2013)). The Advisory Council of Aeronauctical Research in Europe defined

AN US

emissions targets for the year 2020: CO2 reduction about 50 % per passenger kilometer, NOx reduction about 80 % and noise reduction about 50 % compared

to the year 2000 (ACARE (2010)). Similar targets were defined by the US aviation industry (NSaTC (2010)).

To reach these emission targets, lean combustion technologies need to be

M

integrated in future aircraft engines (Wilfert et al. (2007)). For lean combustion high quality fuel atomization is required to ensure a stable combustion in every operating condition during a flight. Prefilming airblast atomizers are used in

ED

aircraft engines to atomize the liquid fuel. Here, a thin liquid fuel film is formed on a prefilmer and disintegrated by coflowing air streams. To effectively control

PT

the atomization an improved understanding of the breakup process is required. A major limitation in understanding these flow systems is a detailed insight into the initial atomization step, the primary breakup. Especially, the impact

CE

of modified geometry and operating parameters is highly demanded to improve the design of combustion chambers. The investigation of the primary breakup is challenging as the geometries

AC

of airblast atomizers are complex and the two-phase flow is characterized by high Weber and Reynolds numbers. To capture the temporal evolution of the liquid film in experimental or numerical investigations high resolutions in time and space are mandatory. In addition, the spatial domain belonging to primary breakup is large compared to the breakup of a liquid jet. Wheras a liquid jet 2

ACCEPTED MANUSCRIPT

is typically injected with a high velocity in a stagnant environment, the velocity of an airblasted sheet is at least one order of magnitude smaller than the coflow air velocity. As a result breakup time and breakup length are extended.

CR IP T

Furthermore, compared to other atomizer designs like flat sheet or jet atomizers, where the dominant instability mechanisms have already been identified

(Marmottant and Villermaux (2004); Fuster et al. (2013)), no such fundamental understanding exists for prefilming airblast atomizers.

Since the idea of an airblast atomizer was first introduced by Lefebvre and

Miller (1966) many experimental investigations have been carried out to under-

AN US

stand the physical processes of the atomization of the liquid film at the atomizing edge. Most of the research did focus on the investigation of the spray charac-

teristic downstream of the atomizer. By applying laser optical measurement systems the key parameters affecting the breakup process were identified. The most pronounced effect on the droplet sizes comes from the mean air velocity within the atomizer. An increase in mean air velocity results in increased aero-

M

dynamic forces that reduce the droplet sizes (Lefebvre (1980); Jasuja (1981); Sattelmayer and Wittig (1986); Aigner and Wittig (1988); Beck et al. (1991);

ED

Bhayaraju et al. (2005)). Additionally, it could be observed that an increase in air pressure reduces the droplet size (Rizkalla and Lefebvre (1975); Lefebvre (1980); Brandt et al. (1996); Bhayaraju et al. (2005); Jasuja (2006)). A reverse

PT

effect on the Sauter Mean Diameter has the increase in the surface tension of the liquid (Lefebvre (1980); Aigner and Wittig (1988); Brandt et al. (1998)). The

CE

liquid viscosity on the other hand has only a minor effect on the droplet size as long as its value is small, which is the case for aero engine fuels (Sattelmayer and Wittig (1986); Jasuja (1981)). A still controversial discussed parameter is

AC

the liquid film thickness. Some researchers observed that an increase in liquid film thickness is accompanied by an increase in the Sauter Mean Diameter (ElShanawany and Lefebvre (1980); Lefebvre (1980); Rizk and Lefebvre (1983)). Other investigations could not identify this behavior and reported that a change in the film thickness has no effect on the droplet size due to a decoupling of the film disintegration process and the film thickness at the atomizing edge. The 3

ACCEPTED MANUSCRIPT

reason is an observed accumulation of the liquid film at the atomizing edge prior to the primary atomization of this accumulation (Sattelmayer and Wittig (1986); Aigner and Wittig (1988); Brandt et al. (1996); Müller et al. (2005);

CR IP T

Gepperth et al. (2013); Bärow et al. (2015)). Therefore, the atomizing edge thickness and air flow structures in the vicinity of the atomizing edge will affect the size of droplets (Gepperth et al. (2012); Braun et al. (2015)).

In order to describe the atomization process analytically, Chaussonnet et al.

(2013) presented an approach that transfers the instabilities observed in co-

flowing atomizer configurations to the breakup of prefilming airblast atomizers.

AN US

To include the accumulation of the liquid at the atomizing edge and its effect on the dominant instability mechanism, the atomization edge thickness is included in the calculation of the transverse instability (Chaussonnet et al. (2013)). However, without detailed numerical investigations it is impossible to identify the responsible atomization instability in prefilming airblast atomizers. To investigate the primary breakup numerically, Direct Numerical Simula-

M

tions (DNS) are a suitable approach as all scales are resolved. The Volume of Fluid (VOF) and the Level Set (LS) are well-established methods to repro-

ED

duce the physics of the immiscible fluids and to transport the interface between them. In several numerical studies the phenomena of airblast atomization are analyzed by computing the breakup of a planar liquid sheet directly injected in

PT

two coflowing air streams. In this configuration no prefilmer is present. Considering this configuration the study by Couderc and Estivalezes (2005) focused on

CE

the validation of global sheet oscillations against experimental results, utilizing a 2-D DNS combined with a LS approach. Villedieu et al. (2013) applied a compressible two-fluid model and compared the oscillation frequency and the

AC

breakup length of the liquid sheet to experimental data. Pringuey and Cant (2014) performed a three-dimensional Large Eddy Simulation (LES) to prove the applicability of the Robust Conservative Level Set (RCLS) method to airblast atomization. This method allows an improved mass conservation of the LS method and is based on the Conservative Level Set by Olsson and Kreiss (2005); Olsson et al. (2007). The injection of a liquid sheet into a stagnant environment 4

ACCEPTED MANUSCRIPT

was investigated by Sander and Weigand (2008) using different injection nozzles and liquid inflow profiles. For the simulations DNS with VOF was combined. Other studies focus on the liquid wall film disturbed by the air stream above,

CR IP T

either with a 2-D RANS solver (Hashmi et al. (2010) or a 3-D DNS combined with a LS-VOF approach (Berlemont et al. (2012)). The 2-D LES simulations

by Volz et al. (2015) focused on the film thickness and were validated against experimental results.

To describe the full atomization process, including primary and secondary breakup, the coupling of an Eulerian multiphase solver with Lagrangian Particle

AN US

Tracking (LPT) is another field of research. Here, small droplets generated by the primary breakup are transferred to a mass point description and tracked by

a Lagrangian solver (Ling and Zaleski (2014); Zuzio et al. (2013)). In contrast to these methods a meshless approach, called Smoothed Particle Hydrodynamics (SPH), is applied by Hoefler et al. (2013); Braun et al. (2015) to simulate primary breakup.

M

Although several numerical and experimental studies have been performed, a combined investigation considering the liquid wall film on top of the prefilmer

ED

and the subsequent disintegration is still a missing link. This paper fills this gap and presents a numerical and experimental investigation of a planar prefilming airblast atomizer, establishing a first basis for identifying the dominating

PT

breakup mechanisms. The study is based on an identical operating point and consistent geometrical conditions of the prefilmer. Further, an identical analy-

CE

sis procedure is applied to the experimental and numerical data, facilitating a reliable comparison of quantitative results, e.g. the breakup length. In order to capture the complex physical processes involved in the primary

AC

breakup of a liquid film in the experiment, sophisticated laser optical diagnostics are used. In combination with a generic atomizer design a highly resolved temporal as well as spacial investigation is possible. For statistically reliable data of droplet sizes and velocities as well as the ligament formation process, an adapted particle and ligament tracking velocimetry setup is used (Müller et al. (2006, 2007)). It was extended by a Depth of Field (DoF) correction algorithm 5

ACCEPTED MANUSCRIPT

to increase the measurement accuracy (Gepperth et al. (2012)). Furthermore, a three dimensional high speed shadowgraphy technique is applied for investigating the various temporal and spatial quantities observed during primary

CR IP T

breakup, leading to a profound understanding of this process. This enables the unique opportunity for accurate quantitative and qualitative comparisons to data obtained by the numerical simulation.

For the simulation the embedded DNS approach is applied as introduced by Sauer (2014) for the primary breakup of airblast atomization. With eDNS a highly resolved 3D-DNS is applied to the breakup region, embedded in a

AN US

LES. High quality boundary conditions are generated by the LES and mapped to the DNS domain. Sauer proved the applicability of the concept using a generic planar prefilming atomizer and six operating points, including cruise and altitude relight conditions.

The paper is organized as follows: the experimental investigation including set-up, diagnostics and post-processing steps are presented followed by the

M

numerical investigation. The embedded DNS methodology and numerical setup are explained. Additionaly, the post-procesing of the DNS data focusing

ED

on an analysis consistent with the experiment, is presented. Furthermore, the selected operating conditions are described. Afterwards, the results of the primary breakup are shown, the following discussion being divided into two sec-

PT

tions. First, the physical phenomena observed both experimentally and numerically are compared considering also the temporal evolution of the liquid sheet.

CE

Second, quantitative results of the liquid sheet and the droplets generated by primary breakup are validated against each other.

AC

2. Experimental Investigation The measurements have been conducted at the spray test rig at the Institut

für Thermische Strömungsmaschinen (ITS). The air is supplied by a radial compressor with a maximum mass flow of 200 g/s. The air mass flow is measured via an air mass flow meter. To ensure a homogeneous velocity field, a mixer, a flow

6

ACCEPTED MANUSCRIPT

straightener and a nozzle are positioned upstream of the test section (Fig. 1). To capture the relevant flow parameters, pressure and temperature probes are connected to the mixing unit. Details on the test rig can be found in Gepperth

CR IP T

et al. (2010); Müller et al. (2006); Müller (2015). 2.1. Test section

In order to enhance the optical accessibility, the test section itself is a two-

dimensional abstraction of an axis-symmetric airblast atomizer (Fig. 1(a)). It consists of four wall segments and the prefilming surface. The air emerging

AN US

from the mixer is split into two air streams that pass the prefilmer on each side. Inside the prefilmer, a cavity used as a continuous feed for the test liquids is located. The liquid is ejected through a set of 50 equidistant distributed holes

with a diameter of 0.5 mm each (Fig. 1(b)). The air shear causes the liquid to form a uniform thin film flowing in the z-direction. It was ensured that no film dryout was occurring during the experiments. The liquid flow through the test

2.2. Diagnostics

M

section is controlled by a mass flow meter/controller.

ED

The experimental observation of the film disintegration process and the measurement of the non-spherical droplets and structures as well as the ligaments

PT

attached to the atomizer is very challenging. For this reason, previous experimental studies focused on the development of an efficient an reliable measurement technique for the observation of the primary breakup (Gepperth et al.

CE

(2012); Müller et al. (2006); Müller (2015)). To derive statistically robust results of the ligament structures and droplet sizes, an efficient algorithm based on

AC

the particle and ligament tracking velocimetry was developed by Müller (2015). The algorithm of Müller (2015) has been extended by a Depth of Field (DoF) correction to further increase the measurement accuracy. Additionally, high speed images are recorded to gain a detailed insight to the

physical processes causing the film disintegration into ligaments and droplets.

7

ACCEPTED MANUSCRIPT

grid

l = 47

21.6 5

46

CR IP T

test section

h = 0.23

8.1

y

prefilmer honeycomb inlet nozzle 65

70.5

AN US

(a) Inlet and generic airblast atomizer

z

x

cavity

liquid inlet

liquid inlet

69.5

ejection holes

x

b = 50 96

ED

z

y

M

wetted area

(b) Prefilmer of the generic airblast atomizer

PT

Figure 1: Experimental setup (all units are represented in mm)

2.2.1. Particle and ligament tracking velocimetry

CE

For the detailed investigation of the liquid film atomization process, a combined method of particle and ligament tracking velocimetry has been used. The technique is based on back light illumination of a measurement volume (shad-

AC

owgraphy) with a double-pulsed laser. The field of view is 12 mm in the zdirection and 16 mm in the y-direction. The spatial resolution is 10 µm. Details on the optical setup and the measurement system can be found in Gepperth et al. (2010, 2012). The key advantage of this technique is that information about the accumulated liquid at the atomizing edge and the generated droplets

8

ACCEPTED MANUSCRIPT

are gathered at the same time. A schematic view of the liquid accumulation at the atomizing edge is shown in Fig. 2. The droplets do not need to be spherical as it is the case with Phase-Doppler-Anemometry (Tropea and Damaschke

CR IP T

(2000); Tropea (2011)). In addition, it is possible to measure the velocity of both, droplets and liquid accumulation at the same time. This is a disadvantage of commercially available shadow sizing systems where only the droplet

velocity can be measured. The information about the liquid structure and the initial particles is of particular interest in the primary atomization region near

the atomizer edge. This helps to link the produced droplets near the atomizer which the droplets are generated.

AN US

edge to the deformation of the liquid accumulation at the atomizing edge by

film flow

air flow

accumulated liquid

y

M

air flow

droplets

z

x

prefilmer

ED

Figure 2: Liquid accumulation at the atomizing edge

The software for the post-processing has previously been developed at ITS

PT

(Kapulla et al. (2008); Müller et al. (2006); Müller (2015)). The post-processing routine can be divided in four major tasks. First, a grey scale analysis of

CE

the recorded double frame images is performed. In this step, a threshold is calculated based on the image intensity distribution. By means of this threshold, the pixel array of the image is than split into background and object. In the

AC

second step, a contouring algorithm is applied. This algorithm detects connected structures based on the grey scale level in step one. It further divides the detected areas into the accumulated liquid at the atomizer edge and droplets. Afterwards, a sphere or ellipse is fitted to the accepted droplets and their size is corrected by a Depth of Field correction (section 2.2.2) based on the work

9

ACCEPTED MANUSCRIPT

of Kim and Kim (1994); Lee and Kim (2004); Castanet et al. (2013). This

0 1 2 3

−5

0 Y [mm]

5

Pstr maxima

AN US

Lstr

CR IP T

Z [mm]

correction leads to a significant increase of the measurement accuracy.

(a) Detected liquid accumulation and ligaments

0 2

6 8

M

Z [mm]

4

ED

10

0 Y [mm]

5

PT

−5

(b) Detected droplets and corresponding velocity com-

CE

ponents

Figure 3: Detected liquid accumulation and droplets after data processing

The third step is the calculation of the droplet velocities in the yz-plane. As a

AC

first estimation a velocity field is calculated based on an ordinary PIV algorithm. Due to the intensity distribution and the huge span of possible droplet sizes, this algorithm produces inaccurate and multiple velocity vectors for each droplet. Therefore, a least square fitting algorithm that uses the estimated velocity field as a ’first guess’ and the detected droplet positions for frame one and two is used

10

ACCEPTED MANUSCRIPT

to calculate the displacement and based on the inter framing time the velocity for each droplet (Müller (2015)). In the last step, the geometrical parameters of the accumulated liquid are

CR IP T

recorded. The blue line in Fig. 3(a) indicates the perimeter of the accumulated liquid, Pstr . The red dots along the perimeter represent the maximum elon-

gation of the liquid accumulation in the streamwise direction. Based on the

position of the atomizing edge and the detected maxima, the breakup length of the ligaments Lstr can be derived. In addition, the detected liquid accumulation

at the atomizing edge of subsequent temporal frames is used to calculate the

AN US

streamwise deformation velocity udef,str of the liquid.

For the present test case, a set of 300 double frame images was post-processed. A processed double frame image is shown in Fig. 3(b), including the raw image, liquid accumulation (blue and red lines), droplets (blue and red ellipses) and the calculated droplet velocities (green arrows). The lengths of the arrows indicate the velocity magnitude and the arrow head points into the direction in which

M

the droplet moves. The atomizing edge is represented by the dashed red line. 2.2.2. Depth of Field (DoF) correction

ED

The back light technique as previously discussed usually does not give a sharp edge of all interesting objects in every part of the image. This is due to

PT

the fact that lenses can only be focused to the so called focal plane. Objects located at exactly this position are accurately captured on the camera chip. If the object moves closer or further away from this focal plane it moves ’out-of-

CE

focus’ and can no longer be detected accurately. The region in which an object can be considered as accurately detected is the depth of field. In an optical

AC

setup this depth of field is dependent on the f-number and the focal distance of the lens, the distance between camera and object as well as the physical pixel size of the CCD chip. In the present study, it was important to capture a wide range of possible droplet sizes, starting from droplets with tenth of microns up to droplets with more than one millimetre. To accurately detect small droplets of less than 60 µm, a large magnification is needed. To achieve this magnification, 11

ACCEPTED MANUSCRIPT

spacer rings were inserted between the lens and the camera. This leads to a

AN US

CR IP T

reduction in the depth of field.

(a) Calibrated droplet diameter as a function (b) Calibrated droplet diameter as a function of the out-of-focus position and reference di- of the out-of-focus position and reference diameter

ameter

2

1.4

M

1.6

x = 0 [mm] x = 1 [mm] x = 2 [mm] x = 4 [mm] x = 8 [mm]

1.2

ED

Dmeas/Dref [mm]

1.8

1

PT

0.8

0.020.1 0.2 0.3 0.4 0.5 Dref [mm]

1

CE

(c) Non-dimensional sizing accuracy for different out-of-focus positions

Figure 4: Calibrated droplet diameter as a function of the out-of-focus position and reference

AC

diameter

On the other hand, the field of view has to have a certain size in order to

be able to track the liquid accumulation for a sufficient length. This leads to a limited resolution of 10 µm per pixel. With such a setup, a constant intensity threshold of 80 %, as it is used in the present investigations, leads to an over estimation of the droplet sizes (Kapulla et al. (2008)). Therefore, a calibration 12

ACCEPTED MANUSCRIPT

is mandatory. In the present study a calibration target was used. Due to the accurate manufacturing process, the spherical dots are highly resolved, both in size and shape. The calibration plate consists of four different areas and covers

CR IP T

dot sizes between 10 µm to 2000 µm. The calibration process itself has to be carried out with the exact same settings (laser power, laser dye, diffuser plate, etc.) that are used for the spray measurements. Therefore, a set of calibration images has been recorded prior to every measurement sequence. The calibration plate was positioned on a 1D traverse system and scanned along the x-direction.

A set of 30 calibration images was recorded for every out-of-focus plane for the

AN US

range of ± 10 mm at every 500 µm. The positive direction refers to a movement

of the calibration plate closer to the camera. A summary of the optical settings can be found in Tab. 1.

Table 1: Characteristic parameters of the optical configuration

Parameter resolution

M

field of view

Unit

Value

mm2

16 × 12

µm/pixel

ED

f-number (f /#)

10 16

focal length

mm

105

calibration planes

mm

−10 to 10

PT

no. of images per plain

30

CE

The same post-processing algorithm as discussed previously for the mea-

surements was used for the analysis of the calibration data sets, but with one difference: in addition to the earlier described detection and ellipse fitting pro-

AC

cess, an additional computation is performed. The detected dot positions are correlated to the known positions of the dots on the calibration plate. The reference diameter of the calibration dot is then compared to the measured dot size. This procedure is repeated for every out-of-focus plane and a diameter calibration matrix (Fig. 4) is derived. In Fig. 4 the color corresponds to the measured

13

ACCEPTED MANUSCRIPT

dot size. The information on the interdependency between the measured and the real dot size can be applied in exactly the same way to the measured droplet sizes during the experiment. It can be seen from Figs. 4(a) and 4(b), that an

CR IP T

almost perfect symmetrical function for the measured dot size over the reference dot size along the x-positions is derived. Details of the difference between measured and detected dot sizes (Fig. 4(c)) reveal the same tendency. Due to the setup and chosen threshold, the diameters are consequently predicted as being

too large. This error increases by moving away from the focal plane. The error in prediction of droplets smaller than 60 µm drastically changes which is likely

AN US

to be due to the optical setup. However, it can be seen that this behaviour sets a limit to the detectable droplet size depending on the out-of-focus position.

Based on these investigations, the diameter correction is a tool that enables the correct diameter measurement of a droplet at an out-of-focus position. However, in a typical setup, as it is used for experimental spray investigations, it is not possible to predict the out-of-focus position of each droplet a priori. Con-

M

sequently, a correction of the droplet diameters cannot be performed based on these predictions. A second parameter is needed to define the out-of-focus posi-

ED

tion for each measured droplet. The intensity gradient at the outer edge of each detected object can usually be used. The mean value of the intensity gradient of every boundary pixel for each dot is calculated and normalized by the median

PT

intensity of the image. This normalized intensity gradient per pixel will be referred to as int/pixel. Using this gradient for every dot on the calibration plate

CE

and in every out-of-focus position, it is possible to create a second calibration matrix for the gradient values (Fig. 5). The color scale refers to the measured mean intensity gradient of each detected dot. Again, the actual gradient values

AC

were only measured for the available dot sizes. The points between the measured dots sizes are linearly interpolated. The gradient and diameter matrixes both show the same behaviour. They are both symmetrical to the yz-plane at x =0 mm. The gradient becomes smaller when moving further away from the focal plane. It can be seen, that for small droplets a peak in the gradient matrix can be seen before the gradient values rapidly decrease. This behaviour is due 14

ACCEPTED MANUSCRIPT

to the experimental setup for the present investigation. The minimum gradient

AN US

CR IP T

value for the evaluation of each diameter class was set to 0.05 int/pixel.

(a) Calibrated normalised intensity gradients (b) Calibrated droplet diameter as a function as a function of the out-of-focus position and of the out-of-focus position and reference direference diameter 2D

ameter 3D

Figure 5: Calibrated normalised intensity gradients as a function of the out-of-focus position

M

and reference diameter

In the measurements for every droplet the normalized intensity gradient and the corresponding measured droplet diameter are known. These two values can

ED

be superposed with the two calibration matrices (Fig. 4(b) and Fig. 5(b)) and the real droplet diameter can be calculated. By applying this methodology,

PT

the calculation can be done without knowing the exact droplet position in the measurement volume. The calculation principle is schematically illustrated in Fig. 6. The interdependency between the measured and the reference diameter

CE

are represented by the grey lines. The numbers in the corresponding fields represent the size of the measured diameter. The black lines and associated values

AC

in the dark fields correspond to the interdependency between the normalized intensity gradient and the real droplet diameter. Both calibration values are symmetrical to the yz-plane at x =0 mm. Therefore, two possible solutions for the real droplet diameter for each combination of measured droplet size and gradient exist. These are given by the two intersection points A and B. Finally, the real droplet diameter in the measurement is calculated by the mean of the 15

ACCEPTED MANUSCRIPT

determined droplet diameter from the two intersection points. The limiting 10

0.05

0.08

intersection A

0.17 0.14 0.11

intersection B 0.08

1

−5

0.20

2

0

0.11 0.14 0.17 0.20

0.5

1

1.5 2 Dref [mm]

2.5

3

AN US

−10

CR IP T

3

x−direction [mm]

2.5

2

1.5

1

0.5

5

Figure 6: Calculation of the real droplet diameter based on the normalized intensity gradient and the measured droplet diameter

parameter for the application of this algorithm is the size of the measurement

M

volume and therefore the lower detectable gradient boundary. In the presented case, this restricts the droplet detection to approximately ± 9 mm in the di-

ED

rection of the optical axis, which has been demonstrated to be well within the range needed for properly studying primary atomization.

PT

2.2.3. Measurement uncertainty The measurement uncertainty has been determined using the approach given by Kline and McClintock (1953). The uncertainty of the mean air velocity can

CE

be estimated to 3 %. The uncertainty of the liquid mass flow is less than 0.5 %.

Due to the enhanced calibration and post-processing procedure, the error in

AC

measuring the droplet size and velocity is estimated to be less than 4 %. In summary, the maximum uncertainties for the measurements are about 7 %.

2.2.4. High speed shadowgraphy The film flow and the primary breakup process have been recorded with two high speed video cameras at 7 kHz at a spatial resolution of 13.89 µm. The field 16

ACCEPTED MANUSCRIPT

of view was set to approximately 14.2 × 14.2mm. Therefore, the utilization of two cameras enables the recording of the liquid film disintegration process simultaneously from two angles shifted by 90◦ . For capturing the high speed breakup

CR IP T

processes the exposure time was reduced to 1 µs, minimizing the blurring effect of fast deforming structures and droplets. For the operating condition a total of

5.000 images are recorded. The measurement volume is illuminated using two 500 W halogen spotlights. The spotlights are mounted opposite of the cameras. 3. Numerical Investigation

AN US

The numerical simulation in this paper is based on the embedded DNS (eDNS) concept. Sauer (2014) introduced this numerical concept for the primary breakup of airblast atomization and proved its applicability by simulating six different operating points. For a detailed validation of the numerical concept with experimental results a further operating point is investigated in the present

M

study. A short overview of the numerical approach is given in the following section, more details can be found in Sauer (2014). The idea of the eDNS concept is to apply a DNS only to a small region

ED

of interest embedded in a larger and coarser LES. For simulating the primary breakup of an airblasted liquid sheet, the DNS domain can be reduced to the primary breakup region, as illustrated in Fig. 7. A key issue of the concept is

PT

the generation of high quality boundary conditions for the embedded domain

AC

CE

by the LES.

Figure 7: The embedded DNS domain for an annular airblast atomizer

17

CR IP T

ACCEPTED MANUSCRIPT

Figure 8: Geometry simplification of the airblast atomizer

The planar configuration, considered here, represents a geometry simplification of the annular atomizer as illustrated in Fig. 8. The application of the

AN US

eDNS approach to the planar airblast atomizer is shown in Fig. 9. The embedded domain comprises a planar prefilmer lip and the atomization region. An

air flow inlet is sited each above and below the prefilmer. The values of the velocity field for the air inlets are generated by a single-phase LES of a turbulent channel. They are extracted on a plane normal to the channel wall at each LES time step and mapped to both air inlets. For the lower air inlet the sign of

M

the y-component is inverted accounting for the boundary layer induced by the bottom side of the prefilmer. As time step and grid size of the LES are coarser

ED

than in the DNS a linear interpolation in time and space is applied to map the data. The data are defined as transient Dirichlet boundary conditions for each individual time step. The turbulent fluctuations are directly transferred to the

AC

CE

PT

DNS domain.

Figure 9: Application of the eDNS concept to a planar airblast atomizer

The application of the transient velocity data to the DNS domain enables an

18

ACCEPTED MANUSCRIPT

investigation of the primary breakup with realistic boundary conditions. Fig. 10 illustrates the evolution of an airblasted liquid film for two different air inflow types: a fully developed turbulent channel flow and a laminar inflow correspondadded in the latter case. Mean velocity profile

3.0 ms

ED

M

4.0 ms

5.0 ms

Time varying profile

AN US

2.0 ms

CR IP T

ing to the mean velocity profile of the same channel flow. No fluctuations are

CE

PT

6.0 ms

Figure 10: Impact of the inlet boundary condition variation on the primary breakup. Gas phase: u = 50 m/s, ρ = 3.9 kg/m3 , ν = 4.7 × 10−6 m2 /s. Liquid phase: u = 0.5 m/s,

AC

ρ = 650 kg/m3 , ν = 2.9 × 10−6 m2 /s, σ = 0.0085 kg/s2 . Prefilmer: h = 50 µm.

The snapshots illustrate that turbulence plays a decisive role in the devel-

opment of the primary breakup regarding the characteristic of the dominating structures. Quantitative parameters as the breakup length or the breakup time are changed by the introduction of turbulence. By applying the embedded DNS 19

ACCEPTED MANUSCRIPT

concept, such influence of the turbulent flow is rationally accounted for and an improved transfer of the results to industrial airblast atomizers is possible. A detailed discussion about the impact of different inflow types on the primary

CR IP T

breakup and the role of turbulence can be found in Sauer (2014). 3.1. Generation of transient inlet boundary conditions for the embedded domain

Due to its close proximity to the prefilmer, the air streams in the embedded domain are mainly affected by wall-related structures. A single-phase LES of a

turbulent channel is selected to generate the suitable velocity data. The quality

AN US

of the generated inflow data influences directly the results of the two-phase DNS simulation.

With a LES the large scales are resolved whereas the small unresolved subgrid scales need to be modeled. This seperation is obtained by filtering of the Navier-Stokes equations, including the continuity Eq. (1) and the momentum Eq. (2), that are formulated with the velocity vector of the resolved scales U ,

M

the gravitational vector g, the kinematic viscosity ν and the kinematic Pressure

ED

P.

(1)

∂U + ∇ · (U U ) = g − ∇P + ∇ · (ν∇U ) ∂t

(2)

PT

∇·U =0

Splitting the convective term of the momentum equation in a resolved and

CE

an unresolved part, (Eq. (3)), the residual stress tensor τ sgs is introduced to

AC

the momentum equation, as formulated in Eq. (4).

U U = U U + τ sgs

(3)

∂U + ∇ · (U U ) = g − ∇P + ∇ · (ν∇U ) − ∇ · τ sgs ∂t

(4)

2 T τ sgs − ksgs I = −νsgs (∇U + (∇U ) ) 3

(5)

20

ACCEPTED MANUSCRIPT

For modeling the subgrid scale stress the Boussinesq approximation is used (Eq. (5)). The turbulent viscosity νsgs and the turbulent kinetic energy ksgs of the subgrid scales are calculated from the One-Equation Eddy model by

CR IP T

Yoshizawa and Horiuti (1985). Fig. 11 illustrates the turbulent channel flow domain which measures 2πδ

in streamwise x-direction, 2δ in wall-normal y-direction and πδ in spanwise z-direction (Moser et al. (1999)). In order that the height of the channel 2δ

M

AN US

corresponds to the prefilmer experiment, δ measures 4 mm.

Figure 11: Geometry and grid of the turbulent channel

ED

A turbulent channel flow is dominated by small wall-related structures. To realize a highly resolved grid in the near-wall region but to limit the general number of grid points a zonal grid is used, refining the domain from the channel

PT

core to the walls. Sauer et al. (2016) compared the results of a coarse, a fine and a zonal LES to the results of a DNS by Moser et al. (1999) and proved the

CE

applicability of the zonal approach. Three different zones are distinguished: a core zone, a wall zone and a sandwiched zone in between them. The number of streamwise and spanwise grid points are doubled from zone to zone, starting

AC

with the core zone. Normal to the walls a simple gradient is applied concentrating more grid points close to the walls. The zonal boundaries are positioned at y + = 200 und y + = 420. Tab. 2 summarizes the non-dimensional grid spacings for each zone, formulated as ∆y + = ∆uτ /ν for the wall-normal distance. Due to + the grid-stretching in y-direction the minimum grid spacing ∆ymin is listed for

21

ACCEPTED MANUSCRIPT

+ the wall and the sandwiched zone, whereas the maximum grid spacing ∆ymax

is listed for the core zone. The grid points in x- and z-direction are equidistant distributed. For the streamwise and spanwise direction periodic boundary con-

CR IP T

ditions are set. The sampling of the velocity data is started once a statistically stationary state of the flow is reached.

The described configuration has also been utilized in the turbulent channel

configuration by Sauer et al. (2016). There the primary breakup of an airblasted

liquid sheet under identical gas phase conditions, but different prefilmer geometry and liquid properties, has been investigated. A detailed discussion about can be found in the cited study.

AN US

the velocity profiles and a validation against DNS results by Iwamoto (2002)

Table 2: Non-dimensional grid spacings for the turbulent channel. Reg = 13, 333, Reτ = 777. + ∆ymin

Wall zone

9.7

4.9

9.2

Sandwiched zone

19.4

17.0

18.5

Core zone

38.8

+ ∆ymax

49.1

∆z +

37.0

ED

M

∆x+

PT

3.2. Two-phase DNS of the primary breakup For simulating the liquid and the gaseous phase the two-phase solver Inter-

CE

FOAM of the opensource code OpenFOAM is used. Spyrou (2011) verified the functionality of the code by computing several test cases for the advection, the surface tension and the buoyancy force. In InterFOAM the Volume-of-Fluid

AC

(VOF) method is implemented to describe the phase interface. The VOF introduces a indicator function α representing the liquid volume

contained in a cell. The density ρ and the viscosity µ in a cell are formulated as a function of the volume fraction (Eq. (6) and Eq. (7)).

ρ = αρl + (1 − α)ρg 22

(6)

ACCEPTED MANUSCRIPT

(7)

µ = αµl + (1 − α)µg

The incompressible Navier-Stokes equations consist of the vectorial momen-

CR IP T

tum advection Eq. (8) under the incompressibility constraint Eq. (9). In order

to find a mathematical description for two-phase flows, the surface tension effect appears as an external force into Eq. (8) ruled by the interface curvature κ. The curvature κ is determined by Eq. (10).

(8)

∇·U =0

(9)

AN US

∂ρU + ∇ · (ρU U ) = ρg − ∇p + ∇ · [µ(∇U + (∇U )T )] + σκ∇α ∂t



∇α |∇α|



M

κ = −∇ ·

∂α + ∇ · (U α) + ∇ · (U r α(1 − α)) = 0 ∂t

(10)

(11)

ED

The transient evolution of the indicator function of an incompressible flow field is described by the advection Eq. (11). The last term on the left side is only active in the proximity of the phase interface and represents an artificial

PT

compression term antagonizing numerical diffusion (Rusche (2002)). No geometrical recontruction of the phase interface is applied. The phase interface is

CE

smeared in a transition region from 0 to 1 and is not known exactly. The surface tension force in Eq. (8) cannot directly be determined. A volume force, that is effective in the transition region, is formulated instead. This method is known

AC

as the Continuum Surface Force (CSF) method by Brackbill et al. (1992). Fig. 12 illustrates the embedded DNS domain. The liquit fuel enters the

domain through a small slit, is transported along the prefilmer to the trailing edge and exposed to high shear forces induced by the coflowing air streams. The liquid accumulates at the trailing edge and is atomized downstream of the trailing edge. 23

CR IP T

ACCEPTED MANUSCRIPT

Figure 12: Computational domain of the planar prefilmer

The prefilmer thickness measures 200 µm and is increased by a factor of four

AN US

compared to the preliminary simulations by Sauer (2014); Sauer et al. (2016). Down to a value of 100 µm the height of the trailing edge has a significant influence on the breakup length and breakup time (Braun et al. (2015)). To

capture the expected breakup length, the computational domain was additionally expanded from 4 mm to 6 mm in the streamwise direction. To keep the DNS domain as small as possible, the prefilmer length only measures 1 mm and

M

is clearly smaller as in the experimental set up. This deviation is accepted, as the studies by Gepperth et al. (2012) have shown, that the prefilmer length has no significant impact on the disintegration process.

ED

The liquid fuel enters the domain with a constant velocity. The faces of the prefilmer are treated as walls with a no-slip condition. An outlet condition is

PT

assigned to the top, the bottom and the outlet of the domain enabling liquid structures to leave the domain. The total pressure is fixed at the outlets and recirculations are prevented by setting inflowing velocity components to zero.

CE

A slip condition is assigned to the lateral faces, the normal component is set to zero and the tangential components are initialized as zero gradients. The

AC

coflowing air next to the prefilmer is directly linked to the turbulent channel flow. The sampled velocity data during the LES is mapped onto the air inlets as transient Dirichlet boundary conditions. For the grid an equidistant grid spacing of 10 µm in all three directions is

realized. For the blocks downstream of the liquid inlet slit and the trailing edge a further refinement by a factor of two in y-direction is performed. Sauer et al. 24

ACCEPTED MANUSCRIPT

(2016) have shown that this grid resolution is suitable to resolve the smallest scales of turbulence, the Kolmogorov length scales lη , regarding the investigated operating point. In their study, the embedded DNS concept has been utilized

CR IP T

to simulate the primary breakup in the same turbulent channel flow (Reg = 13, 333). The prefilmer thickness and the liquid properties were modified. The Kolmogorov length scale was found to be lη = 12 µm and the required minimum

grid spacing was identified as ∆xmin = 25 µm following Eq. (12) by Pope (2000).

The relation between the minimum required grid spacing in physical space and

AN US

the Kolmogorov scales is quantified as ∆xmin π ≈ 2.1. = lη 1.5

(12)

A further aspect in the grid resolution assessment is the representation of small liquid structures and droplets, produced by the primary breakup. The findings by Sauer et al. (2016) show that an resolution of the entire liquid mass is hard to achieve due to the amount of small droplets. 80 % of the liquid

M

mass have been resolved in their simulation utilizing the above mentioned grid resolution. A detailed discussion about the droplet representation on different

ED

grid sizes (20 µm and 10 µm) can be found in the cited study. For the operating point, investigated in this paper, the liquid structure resolution is discussed in section 6.2 and 6.3, directly linked to the quantitative results.

PT

The total number of cells amounts 84.41 millions. 14 ms physical time with a time step of 50 ns has been simulated. The highly resolved DNS was performed

CE

on 720 processors. A total of 540000 cpuh were consumed.

AC

3.3. Processing of DNS data Analogously to the experimental data processing, the analysis of the com-

puted results is splitted into the analysis of the accumulated liquid at the trailing edge and the analysis of droplets. For determining the number of droplets and their properties, a sampling

plane at the outlet is defined. According to a minimum volume fraction α in each cell, a connected liquid structure is detected by using an approach of Connected 25

ACCEPTED MANUSCRIPT

Component Labeling (CCL). The surface area of each droplet is quantified by the sum of all cells i containing liquid with the corresponding volume fraction αi and the grid sizes ∆yi and ∆zi (Eq. (13)). The droplet diameter is derived

CR IP T

by assuming a circular area. Velocity data is extracted at each cell belonging to a droplet. The components of the droplet velocity Ud are further calculated as the mean value over the area of the droplet (Eq. (14)). This post-processing

routine is inspired by the algorithm presented by Grosshans and Fuchs (2011) and Herbert et al. (2008). Ad =

n X

αi ∆yi ∆zi

AN US

i=1

Pn i=1 αi Ui Ud = P n i=1 αi

(13)

(14)

The analysis of the accumulated liquid at the trailing edge is challenging. An analysis at a fixed position, similar to the droplets, is not suitable as the

M

liquid film oscillates and changes its shape and position from time step to time step. To determine characteristic sizes, like the breakup length and the streamwise deformation velocity, the position of the phase interface needs to be known.

ED

Therefore an post-processing algorithm has been developed, facilitating an automated analysis of the continuous liquid film.

PT

The procedure is illustrated in Fig. 13. Based on a predefined α-threshold the phase (liquid or gas) in a grid cell is determined. Corresponding to the identified phase each cell receives the label 1 (liquid) or 0 (gas). A view from

CE

the top, analogous to the shadowgraphy images, is generated by summing up the labels in x-direction. A sum exceeding 1 signifies that at this position a liquid phase would be observed from the top. A CCL algorithm is applied,

AC

identifying all liquid structures. A connectivity of 8 is utilized, checking for liquid in each neighboring cell. As only the accumulated liquid at the trailing edge is of interest, any other liquid structure, detected by the CCL, is deleted. Finally, the interface of the accumulated liquid is identified. The result is a one-dimensional plot of the phase interface, that enables an easy and efficient

26

M

AN US

CR IP T

ACCEPTED MANUSCRIPT

ED

Figure 13: Reducing three-dimensional DNS data to a plot of the interface

analysis compared to the original three-dimensional DNS data. The described processing is applied for every time step. The subsequent procedure is reduced

PT

to the analysis of the recognized phase interface, similar to the experimental data processing. This way an identical definition of the breakup length and the deformation velocity for the experimental and numerical investigations is

CE

possible.

Reducing the huge amount of DNS data to single plots of the phase-interface

AC

enables a fast and efficient analysis. But in contrast, two disadvantages have to be mentioned: first, superimposed structures as seen from the top can not be distinguished and thus an overestimation of the liquid sheet expansion into the atomization region is possible. Secondly, the velocity data calculated by the DNS is not used. Both aspects could be avoided by identifying the accumulated liquid in the original DNS domain and extracting the velocity data in each cell 27

ACCEPTED MANUSCRIPT

belonging to the liquid sheet or its interface. However, due to the focus of this paper, the DNS data is here reduced, facilitating a more reliable comparison

CR IP T

against the experimental results. 4. Operating Conditions

Real aircraft engine conditions are characterized by high Reynolds and Weber numbers providing highly turbulent flow conditions and finely atomized droplets. A detailed investigation requires a high resolution in time and space,

challenging for both simulations and experiments. For the present study a mod-

AN US

erate operating point with a comparatively low Reynolds- and Weber number

has been selected. It corresponds to altitude relight conditions and enables an adequate resolution of the breakup events in both investigations. It is therefore suitable for a reliable comparison of the numerical and experimental results. The operating conditions are summarized in Tab. 3. The gas Reynolds num-

M

ber Reg is formulated with the half height of the channel δ as Reg = ug δ/νg . Here, ug and νg describe the velocity and the kinematic viscosity of the gas 2

phase. The film weber number, W er,ht = ρg (ug − ul ) ht /σ, considers the in-

ED

fluence of the density of the gas phase ρg , the velocity difference between gas and liquid as well as the surface tension σ. As characteristic length the prefilmer trailing edge thickness is utilized due to its impact on the breakup length

PT

identified by Gepperth et al. (2012). To classify the atomization process the momentum flux ratio between gas and liquid phase (M = ρg u2g /ρl u2l ) is also

CE

specified.

In the afore mentioned study by Gepperth et al. (2012) the influence of dif-

ferent properties of several liquids has been investigated besides the influence of

AC

geometrical parameters like prefilmer length and trailing edge thickness. From the investigated liquids Shellsoll D70 with a low surface tension at ambient conditions has been selected. The surface tension plays a decisive role as it antagonizes the disintegration. When the surface tension force is low, the interface behaves less stable and breakup events occur faster. Therefore the size of the

28

ACCEPTED MANUSCRIPT

Table 3: Operating conditions

Unit

Value

Reg

13,333

W er,hl

21.4

ρl /ρg

636.4

M

16.0

Gas properties m/s

ug

kg/m

1.2

AN US

ρg

50 3

CR IP T

Parameter

m /s

1.50e-05

m/s2

0.5

kg/m3

770

m2 /s

2.03e-06

2

νg

Liquid properties ul ρl νl

kg/s

0.0275

mm2 /s

50

ED

V˙ /b

M

σ

2

numerical domain can be kept small in the present investigation. In addition

PT

the physical properties of Shellsoll D70 are close to those of aviation fuels. The liquid volume flow rate per unit length V˙ /b is kept constant at 50 mm2 /s

CE

during both the experiment and the simulation. At the liquid inlet slit of the DNS domain the velocity is set to 0.5 m/s. In contrast, the bulk air velocity of

AC

the air is 50 m/s. 5. Qualitative Results In Fig. 14 a snapshot of the disintegrating liquid film obtained by the shad-

owgraphy is shown. A red rectangle illustrates the cross section of the DNS domain in comparison to the experimental field of view. Due to the numerical 29

ACCEPTED MANUSCRIPT

effort, the DNS domain is smaller and actually reduced to the primary breakup region. The domain covers 1 mm upstream and 5 mm downstream of the prefilmer trailing edge in normal direction. In parallel to the trailing edge the

CR IP T

domain measures 4 mm. In both investigations the following breakup process is observed: waves are formed on top of the prefilmer, transporting the liquid to the prefilmer trailing edge, where a reservoir is formed. Out of this reser-

voir elongated structures (ligaments) as well as bag like structures are formed, grow in size and finally break up into droplets. The droplets differ in their size,

small as well as big bulgy droplets are generated. During the atomization no

PT

ED

M

AN US

dewetting of the prefilmer is observed.

Figure 14: Experimental field of view and the used image section of the size of the DNS

CE

domain for the comparison of the breakup phenomena

To visualize the breakup process, two different snapshot sequences are shown

in Fig. 15 and Fig. 16. The side- and topview of the experiment and the sim-

AC

ulation are shown. All image sections correspond to the dimensions of the numerical domain. The time lag between two numerical snapshots is 0.3 ms, which corresponds to every second high speed shadowgraphy image (frame rate 7 kHz). The time stamp of the first snapshot in Fig. 15 is set to 0 ms and serves as reference time.

30

ACCEPTED MANUSCRIPT

In the first sequence the formation of ligaments, arising from the liquid reservoir at the trailing edge, is visualized. The surface tension force can not counteract against the disintegration of the accumulated liquid so that the liquid

CR IP T

starts to break up. Initial holes are formed and grow in time. Separated ligaments are left over, that are elongated deeper into the atomization region until parts of the ligaments are separated. In the experiment as well as in the simula-

tion varying ligament dimensions have been observed. Especially the maximum

length of single ligaments differs, a stable breakup length could not be identified. From the sideview of the numerical snapshots weak oscillations around

AN US

the prefilmer (on the x-axis) are noticed. Thereby, a synchronous flapping of all ligaments along the prefilmer is not observed. This becomes apparent in the experimental sideview images, that capture a higher number of breakup events related to the increased experimental prefilmer width. The large variation in the ligament flapping leads to a strong overlapping and at least to a reduced sharpness of the sideview images.

M

In the second sequence, starting at 2.9 ms, the ligaments have already been torn away. Out of the liquid reservoir a bag is formed by the gaseous flow,

ED

that increases particularly fast and finally bursts. The rim of the bubble-like structure is more stable and remains attached to the reservoir for a short time, before it disintegrates into ligaments or bigger droplets. The generation of

PT

bubbles occurs fast compared to the ligament formation in Fig. 15. From 2.9 ms to 3.1 ms the bag observed in the numerical snapshots has already been bursted

CE

and a second bag has started to develop close to the trailing edge. The same dynamic is noticed in the experiment: at 2.86 ms only the rim of the bag is

visual, at 3.43 ms a new bag is generated, that has already been bursted at

AC

3.72 ms. The disintegration of the bags or bubble-like structures results in the generation of multiple droplets and is more intensive compared to the ligament breakup. In conjunction with the formation of bags, strong oscillations around the prefilmer are visible. The investigations showed that the bags are formed alternating up- and downwards, according to an oscillation to the right and to the left in the sideview snapshots. However, similar to the ligaments this 31

ACCEPTED MANUSCRIPT

appears to be a stochastic process and flapping of all generated bags to the same side of the prefilmer with the same frequency is not observed. It should be noted that no continuous liquid sheet is developed behind the

CR IP T

prefilmer. The bags and ligaments arise directly from the accumulated liquid at the trailing edge, flapping not in phase. As a consequence it is hard to identify

a global oscillation frequency as it is typical specified for flat sheet atomizer

designs. In addition, the bags and ligaments along the prefilmer do not start to

develop at the same time, whereby the spanwise spacing between ligaments or bags is changing permanently. The inter-ligament distance is strongly affected

AN US

by separating or new arising ligaments and bags at varying positions.

With a momentum flux ratio of 16 the primary breakup of the analyzed operating point is located inside the torn sheet breakup regime according to the breakup regime classification by Fernandez (2010). The classification is oriented to the morphological aspects of the breakup instead of instability mechanisms. Fernandez specified a momentum flux ratio between 5 and 22 for the torn-sheet

M

breakup. Typically, the liquid sheet is eventually torn away due to the shear effect by the gaseous flow. The generation of longitudinal streamwise ligaments,

ED

irregular shaped, and the formation of big bulgy droplets is also significant for the torn-sheet regime. Referring to the images in Fig. 15 and Fig. 16 a wide

AC

CE

PT

range of droplet sizes, including big bulgy droplets are generated.

32

ACCEPTED MANUSCRIPT

x

y

z

y

z

x

z

0.0 ms

0.3 ms

M

0.29 ms

AN US

CR IP T

0.0 ms

z

0.6 ms

0.86 ms

0.9 ms

AC

CE

PT

ED

0.57 ms

33 Figure 15: Disintegration sequence at the atomizing edge with ligament formation. From left to right: Sideview experiment, topview experiment, topview simulation, sideview simulation

ACCEPTED MANUSCRIPT

x

y

z

y

z

z

2.9 ms

CR IP T

2.86 ms

AN US

3.1 ms

M

3.14 ms

x

z

3.4 ms

3.72 ms

3.7 ms

AC

CE

PT

ED

3.43 ms

34

Figure 16: Disintegration sequence at the atomizing edge with the formation of bag like structures. From left to right: Sideview experiment, topview experiment, topview simulation, sideview simulation

ACCEPTED MANUSCRIPT

6. Quantitative Results Quantitative data are obtained in order to compare the numerical and experimental results. The accumulated liquid film at the trailing edge is characterized

CR IP T

by its breakup length and its streamwise deformation velocity. The analysis pro-

cedure to extract those quantities is presented, providing a consistent definition for the numerical and experimental results. As the numerical results are linked

to a predefined threshold for the liquid volume fraction α (see section 3), the influence of this parameter on the numerical results is additionally discussed. Finally the droplet sizes and droplet velocities are compared. In addition to

AN US

their mean values the distributions of all compared quantities are presented. 6.1. Characterization of the accumulated liquid

The analysis of the accumulated liquid is based on the phase interface. For the experimental investigation the analysis procedure is included in the software by Müller (2015), utilized for the post-processing of the shadowgraphy

M

images. The analysis of the numerical results has been implemented in the post-processing routine introduced in section 3.3. Both algorithms follow the

ED

same procedure to determine the breakup length and the streamwise deformation velocity as described in the following section. The expansion of the liquid film into the atomization region is quantified by

PT

the length of ligaments, arising from the accumulated liquid. A single ligament length Lstr is defined as the distance between the ligament peak and the trailing

CE

edge of the prefilmer. The mean value of all ligament lengths is defined as the breakup length Lbu . Ligaments are identified as maxima of the phase interface plot. A central

AC

aspect is a well-defined choice of the detected maxima to suppress repeated counting of ligaments with a wrinkled contour. As a criterium a minimum distance between adjacent minima and maxima of 50 µm is predefined to obtain a more global behaviour, as illustrated in Fig. 17. The choice of the minimum distance is always a trade-off, as also small ligaments close to the trailing edge should be detected. 35

1

1

2

2

3

3

4

4

5

5

6 −2

−1

0 [mm]

1

2

6 −2

−1

0 [mm]

CR IP T

0

[mm]

0

1

2

AN US

[mm]

ACCEPTED MANUSCRIPT

(a)

(b)

Figure 17: Peak detection for the phase interface (a) and the effect of a minimum distance of 50 µm (b) for the analysed operating point

The breakup length is obtained as a mean value of several snapshots, using

M

the current length of each ligament. This means a significant modification for the analysis of the numerical results in relation to previous studies by Sauer

ED

(2014), where the breakup length was defined as the maximum ligament length taken from a series of time steps. The analysis procedure has been started after the first breakup event occured.

PT

In case of the simulation ligaments attached to the sides of the numerical domain are not included in the analysis. Due to the slip condition, mentioned in section 3.2, these ligaments are prevented from breaking through the sides.

CE

In the simulation no significant deviation in the ligament development has been observed for ligaments in the vicinity of the sides or in the center of the domain,

AC

as the dominating expansion is in the streamwise direction. However, to avoid any errors, these ligaments are neglected. To determine the streamwise velocity of the continous liquid film in the

experiment the displacement of the phase interface between two timesteps is used. This approach is also adapted for the numerical results, providing an identical definition of the deformation velocity. The unique assignement of a 36

ACCEPTED MANUSCRIPT

specific interface point at two consecutive timesteps t1 and t2 is difficult due to variations in the wrikled contour or the seperation of liquid structures. In order to simplify the calculation, the determination of the displacement of each

CR IP T

interface point is replaced by utilizing only the most remote interface point zmax per y-position at t1 and t2 . The streamwise deformation velocity at position yi is then calculated as udef,i = zmax,yi (t2 ) − zmax,yi (t1 )/(t2 − t1 ).

Fig. 18 illustrates the phase interface contour at two consecutive timesteps

and a cumulative distribution of the calculated velocities. Negative velocities indicate the separation of liquid structures (E1), whereas very high velocities

AN US

indicate the expansion or movement of a ligament to the side, leading to an overestimation of the displacement at a close-by y-position (E2). The maximum velocity per frame or timestep is well represented by the 90 %-quantile, its mean

value of all timesteps provides the characteristic deformation velocity udef,str of

E2 Udef,i [m/s]

ED

E2

E1 N/N total

y [mm]

(a)

(b)

CE

PT

z [mm]

E1

M

the investigated operating point.

Figure 18: Time displacement of the phase interface and ambiguities by its determination

AC

6.2. Breakup length and deformation velocity The breakup length Lbu and the deformation velocity udef,str determined

from the experiment and the DNS are summarized in Tab. 4. For their determination the phase interface position is required, that is found by a predefined liquid volume fraction (α = 0.5) in case of the simulation. In oder to study the 37

ACCEPTED MANUSCRIPT

sensitivity of the numerical results against α, additional thresholds have been utilized in the post-processing routine described in section 3.3. In Tab. 4 the numerical results for α = 0.5 and α = 0.01 are shown and compared to the

CR IP T

experimental results. Table 4: Characteristic values for the accumulated liquid

α = 0.5

α = 0.01

Lbu [mm]

3.2

2.2

2.7

udef,str [m/s]

15.7

8.2

14.4

fbu [1/s]

4906.3

3727.3

5259.3

AN US

Experiment

The breakup lengths of the DNS are a little smaller, albeit all breakup lengths are in the same order. For a better understanding of the determined

M

values, the distribution of the ligament lengths is provided in Fig. 19. In the experiment a small amount of ligaments greater than the DNS domain (Lbu > 5 mm) are observed, leading to a greater breakup length. In the simulation the

ED

accumulated liquid reaches the outlet of the computational domain at no time step, the maximum ligament length measures 4.8 mm. Nevertheless, a reliable prediction of a maximum ligament length is difficult, due to the size of the

PT

numerical domain in streamwise direction. The long ligaments detected in the experiment are single structures, typically with a very thin connection to the

CE

liquid film. They appear occasional and can be detected by shadowgraphy due to a greater time period of recording. It is based on 300 frames with a time span of 100 ms between each frame. In contrast to that, the numerical results

AC

of 190 time steps with a time span of 20 µs are analyzed. Due to the numerical effort, the physical simulation time is kept short and a time span corresponding to the experiment is not feasible. The variation of the numerical threshold for the volume fraction implies

a displacement of the interface and has an effect on the numerical results by

38

ACCEPTED MANUSCRIPT

0.3

0.2

0.1

0 0

5 Lstr [mm]

0.2

0.1

0 0

10

Experiment Sim. α=0.01

CR IP T

Experiment Sim. α=0.5 N/Ntotal

N/Ntotal

0.3

5 Lstr [mm]

10

Figure 19: Distribution of the ligament length 0.2

0.15

N/Ntotal

0.1 0.05

Experiment Sim. α=0.01

AN US

total

0.15 N/N

0.2

Experiment Sim. α=0.5

0.1

0.05

0 0

20 40 Udef,str [m/s]

60

0 0

20 40 Udef,str [m/s]

60

M

Figure 20: Distribution of the deformation velocity

principle. As the grid resolution is small compared to ligament lengths the

ED

influence by the interface smearing is minor. Critical is the point of separation, when the connection between a ligament and the remaining accumulated liquid

PT

becomes very thin and is finally below the grid resolution. If α is reduced, the separating ligament is seen attached a few more timesteps, while it is further expanded in the streamwise direction. This effect is seen in the ligament length

CE

distribution for α = 0.01. The threshold is extremely low and illustrated as a lower bound. However, as illustrated in Fig. 21, the overall impact on the characteristic value for the breakup length is small. The deviation from Lbu (α =

AC

0.5) is smaller than ±7 % in the range of 0.1 < α < 0.9 and comparable to the experimental measurement uncertainty.

As specified in Tab. 4, a twice as high deformation velocity is measured

in the experiment in comparison to the DNS, if α = 0.5 is predefined for the phase interface. In contrast to the breakup length, a significant impact of the α39

ACCEPTED MANUSCRIPT

16

L [mm] bu

U

[m/s]

def

8

4

0

0.2

0.4

α

0.6

0.8

CR IP T

12

AN US

Figure 21: Sensitivity of breakup length and deformation velocity dependent on the liquid

M

fraction α

ED

(a)

(b)

(c)

Figure 22: Representation of bubbles in the experiment (a) and the simulation with α=0.5

PT

(b) or α=0.05 (c)

threshold on the deformation velocity has been noticed, analyzing the numerical

CE

data. Reducing this threshold, the velocity distribution expands to greater velocities, assimilating the experimental distribution as shown in Fig. 20. The deformation velocity at several α-values is shown in Fig. 21. A moderate change

AC

for the deformation velocity is observed down to α = 0.2. Below this value, the velocity rises steeply, reaching a value of nearly 15 m/s at a liquid volume fraction of 1 %. It is presumed that the influence on the velocity can be attributed to an insuf-

ficient resolution of bubbles, arising from the accumulated liquid (see section 5).

40

ACCEPTED MANUSCRIPT

A typical bubble, observed during the experiment, is shown in Fig. 22(a). It is characterized by a thicker rim and a bag formed by a very thin liquid film. Typically, these bags rise faster than the rim into the atomization region and reach

CR IP T

high velocities when they burst. In the simulation bag-like structures are also observed, but as the bags grow and the liquid sheet becomes thinner, a suitable resolution is hard to achieve. This is visualized in Fig. 22(b) and Fig. 22(c).

Considering a low threshold of α, the liquid film of a bubble is visible and higher velocities are observed.

As discussed in section 5 a global oscillation frequency is hard to determine

AN US

as no continuous flapping sheet exists behind the trailing edge. As an alternative

the mean breakup frequency is calculated as a measure for the time distance between consecutive atomization events. It is directly obtained from the deformation velocity and the breakup length as fbu = udef /Lbu . For the analyzed operating point the frequency is in the range of 5000 1/s as specified in Tab. 4. Due to the variation of the deformation velocity for the DNS data only a range

M

for the breakup frequency can be determined.

ED

6.3. Droplets

For the design of a combustion chamber the properties of the droplets generated by the airblast nozzle are a valuable information. The droplet size distribu-

PT

tions from the experiment and the simulation are shown in Fig. 23. The droplet diameters are obtained 5 mm downstream of the trailing edge, corresponding to the outlet position of the numerical domain. This analysis position differs

CE

from preliminary experimental studies by Gepperth et al. (2012), in which the entire field of view has been taken as a basis. The adaption improves the com-

AC

parability of the experimental and numerical results. Similar to the liquid film analysis two sets of numerical droplet data are compared to the experimental results: droplet data regarding a predefined α-threshold of 0.5 and a minor one (α = 0.03). In all three cases droplets with a diamteter up to 300 µm are observed (Fig. 23). For the numerical results a high number of small droplets in the 41

ACCEPTED MANUSCRIPT

range of a few micrometers is noticeable. For α = 0.03 the shape of the distribution differs strongly from the experimental results. Sauer et al. (2016) defined the liquid detected by this low threshold as the atomized mass that is not re-

CR IP T

solved by the grid. This definition promotes the localization of small subgrid droplets, that are also not detectable by the experimental resolution. Experiment Sim. α=0.5 Sim. α=0.03

0.03 0.02 0.01 0 0

AN US

PDF [1/µm]

0.04

100

200 dd [µm]

300

Figure 23: Droplet size distribution: Comparison between experiment and simulation

The knowledge of the smallest droplets generated by primary breakup is

M

a fundamental problem caused by the resolution limit of numerical as well as experimental investigations. From the numerical simulation a determination of the liquid mass, that is not resolved by the numerical grid, is possible. As the

ED

VOF approach is mass conservative, the choice of the α-criterion defines how many mass is lost in the process. Each cell containing at least a volume fraction of α is taken into account for the droplet analysis. Droplets determined by

PT

α = 0.5 are defined as resolved droplets in this paper. For the quantification of the resolved liquid mass the droplet mass md at the outflow of the DNS

CE

domain is studied for varying α-values. In addition, the corresponding number of droplets Nd , passing through the outflow plane, is determined. The results

AC

are visualized in Fig. 24, normalized by their respective value taken at α = 0.01. The evaluation of the liquid mass depending on α shows that 80 % of the

total liquid mass are represented for α = 0.5. This mass is carried by 20 % of the liquid structures. It becomes evident that the large and mass carrying structures are resolved by the numerical grid, but that a resolution of the total liquid mass due to the huge amount of fine droplets is hard to achieve. Fig. 23,

42

ACCEPTED MANUSCRIPT

m /m d

d,total

N /N d

d,theo

1

0

0.2

0.4

α

0.6

CR IP T

0.5

0.8

Figure 24: Small structure properties extracted at the eDNS outflow: droplet mass and number of droplets

which depicts the droplet distribution for α = 0.03, includes these fine droplets,

AN US

not resolved, but estimated as atomized mass.

Comparing the numerical and experimental droplet data the definition of a minimum droplet size, detectable in both investigations, is useful. In the experiment the lower limit of a droplet diameter is 20 µm, assuming a resolution of a single droplet by two pixels. Neglecting all droplets with a diameter smaller

M

than 20 µm for the numerical data, the droplet distributions in Fig 25(a) are obtained. A good agreement of numerical and experimental results is observed. As a resolution of 10 µm is already realized by the numerical grid, the choice of

ED

α has no specific influence.

0.03 0.02

0 0

Experiment Sim. α=0.5

0.15

0.01

CE AC

0.2

PDF [s/m]

Experiment Sim. α=0.5 Sim. α=0.03

PT

PDF [1/µm]

0.04

0.1 0.05

100

200 d [µm]

0 0

300

d

20

u [m/s]

40

60

x

(a)

(b)

Figure 25: Droplet distribution for diameter and streamwise velocity: Comparison between experiment and simulation

The droplet velocity distributions in the streamwise direction are shown in Fig. 25(b). The mean values of both investigations are of the same order, mea43

ACCEPTED MANUSCRIPT

suring 14.5 m/s for the experiment and 10.4 m/s for the simulation. Nevertheless a significant amount of droplets with a velocity above 15 m/s is observed in the experiment, leading to a wider distribution and a higher mean velocity as in

CR IP T

the simulation. It is presumed that the insufficient numerical resolution of the bubbles leads to an underestimation of the droplet velocity in the vicinity of the atomizing edge. As the fast development of a bubble is not predicted completely (see section 6.2), a correct prediction of the droplet acceleration, induced by the bursting of bubbles, is also prevented.

80

40

150

300 dd [µm]

450

M

0 0

Experiment Simulation

AN US

uz [m/s]

120

Figure 26: Scatter plot for the droplet diameter dd and the streamwise droplet velocity uz .

ED

A deeper insight into the droplet properties is obtained by the scatter plot of the droplet diameter and velocity in Fig. 26. Raw data is shown, including the small droplets (dd < 20 µm) in case of the simulation. In addition, high droplet

PT

velocities in the experimental data (50 - 115 m/s) that are found due to errors in the droplet detection algorithm, are not filtered. These high values correspond

CE

to a false detection of the same droplet in two consecutive images, leading to an overestimation in the droplet velocity. In the scatter plot a wider spread of the experimental droplet properties can be observed. This is due to a greater time

AC

period for the sampling, and therefore, a higher amount of samples compared to the simulation. Single droplets with a velocity up to 50 m/s are recorded as well as droplet diameters up to 500 µm. These droplets occur occasionally and

have no significant influence on the number based distributions. In contrast, the Sauter mean diameter (as a volume based quantity) is affected even by a small number of larger droplets. In the present study it measures 154.8 µm for the 44

ACCEPTED MANUSCRIPT

experiment and 128.1 µm for the simulation. To improve the statistics regarding very large droplets in the simulation, a greater simulation time is required. Nevertheless, comparing the number based distributions, a good agreement has

CR IP T

been identified. 7. Conclusions

In this paper a detailed comparison of numerical and experimental results of the primary breakup of a prefilming airblast atomizer was presented. Particle and ligament tracking velocimetry based on back light illumination was applied

AN US

in the experiment, whereas the embedded DNS methodology was utilized in the simulation. An operating point in the range of altitude relight conditions

has been studied. A detailed discussion of the observed breakup phenomena and quantitative results of the breakup length, the deformation velocity of the liquid accumulation, droplet sizes and droplet velocities demonstrated the com-

M

parability of the results. To ensure a consistent definition of the characteristic dimensions in case of the liquid accumulation, the three-dimensional DNS data was reduced to the topview corresponding to the shadowgraphy images. There-

ED

fore, the same analysis procedure could be applied to the experimental and the numerical data. Both investigation methods provided a similar prediction of the primary breakup, but an underestimation of large ligaments and very large

PT

droplets could be identified for the simulation. As observed in the experiment, these liquid structures appear occasionally and have a minor effect regarding

CE

the total number of droplets or ligaments. The comparison clearly illustrates the challenges for the spatial resolution of a thin liquid film in the numerical investigation. Bubbles arising from the liquid reservoir at the trailing edge were

AC

insufficently resolved as they grow and the liquid film becomes thinner. A further grid refinement down to 5 µm is in the focus of a current study. For the comparison of the droplet data a minimum size of 20 µm was specified, cor-

responding to the resolution of the experiment. Based on this value a good agreement between numerical and experimental droplet data was found. The

45

ACCEPTED MANUSCRIPT

results confirm the applicability of both methods, offering a detailed insight in the phenomena of primary breakup during airblast atomization.

CR IP T

Acknowledgment The authors would like to thank the German Research Foundation (DFG) for the financial support under the contract no. JA 544/39-1 and GRK 1344.

The authors gratefully acknowledge the Gauss Centre for Supercomputing e.V. for funding this project by providing computing time on the GSC Super-

AN US

computer SuperMUC at Leibniz Supercomputing Centre. References

ACARE, 2010. Aeronautics and air transport: Beyond vision 2020 (towards 2050). Technical Report.

M

Aigner, M.M., Wittig, S., 1988. Swirl and counterswirl effects in prefilming airblast atomizers. Journal of Engineering for Gas Turbines and Power 110,

ED

105–110.

Airbus, 2013. Airbus global market forecast 2013-2032. Technical Report. Bärow, E., Gepperth, S., Koch, R., Bauer, H.J., 2015. Effect of the precessing

PT

vortex core on primary atomization. Zeitschrift für Physikalische Chemie 229, 909–929.

CE

Beck, J.E., Lefebvre, A.H., Koblish, T.R., 1991. Airblast atomization at conditions of low air velocity. Journal of Propulsion and Power 7, 207–212.

AC

Berlemont, A., Bouali, Z., Cousin, J., Desjonquères, P., M.Doring, Ménard, T., Noël, E., 2012. Simulation of liquid/gas interface break-up with a coupled level set/vof/ghost fluid method, in: Seventh International Conference on Computational Fluid Dynamics (ICCFD7), Big Island, Hawaii.

46

ACCEPTED MANUSCRIPT

Bhayaraju, U.C., Giuliani, F., Hassa, C., 2005. Planar liquid sheet breakup of pre- and nonprefilming airblast atomisers at elevated ambient air pressure, in: ILASS Europe, 20th Annual Conference on Liquid Atomization and Spray

CR IP T

Systems, Orleans, France. Boeing, 2013. Current market outlook 2013-2032. Technical Report.

Brackbill, J., Kothe, D., Zemach, C., 1992. A continuum method for modelling surface tension. Journal of Computational Physics 100, 335–354.

Brandt, M., Gugel, K.O., Hassa, C., 1996. Experimental investigation of the

AN US

liquid fuel evaporation in a premix duct for lean premixed and prevapor-

ized combustion, in: 1996 ASME: International Gas Turbine and Aeroengine Congress & Exhibition, Birmingham, UK.

Brandt, M., Rachner, M., Schmitz, G., 1998. An experimental and numerical study of kerosine spray evaporation in a premix duct for gas turbine combus-

M

tors at high pressure. Combustion Science and Technology 138, 313–348. Braun, S., Wieth, L., Koch, R., Bauer, H.J., 2015. Influence of trailing edge

ED

height on primary atomization: numerical studies applying the smoothed particle hydrodynamics (sph) method, in: ICLASS, 13th Triennial International Conference on Liquid Atomization and Spray Systems, Tainan, Taiwan.

PT

Castanet, G., Dunand, P., Caballina, O., Lemoine, F., 2013. High-speed shadow imagery to characterize the size and velocity of the secondary droplets pro-

CE

duced by drop impacts onto a heated surface. Experiments in Fluids 54, 1–17.

AC

Chaussonnet, G., Riber, E., Vermorel, O., Cuenot, B., Gepperth, S., Koch, R., 2013. Large eddy simulation of a prefilming airblast atomizer, in: ILASS Europe, 25th Annual Conference on Liquid Atomization and Spray Systems, Crete (Chania), Greece.

Couderc, F., Estivalezes, J.L., 2005. Direct numerical simulation of plane liquid sheet atomization : a parametric study, in: Proceedings of ILASS Americas, 47

ACCEPTED MANUSCRIPT

18th Annual Conference on Liquid Atomization and Spray Systems, Irvine, CA, USA. El-Shanawany, M.S., Lefebvre, A.H., 1980. Airblast atomization: The effect of

CR IP T

linear scale on mean drop size. Journal of Energy 4, 184–189.

Fernandez, V.G., 2010. Experimental Study of a Liquid Sheet Disintegration in a High Pressure Environment. Ph.D. thesis. Université de Toulouse.

Fuster, D., Matas, J.P., Marty, S., Popinet, S., Hoepffner, J., Cartellier, A.,

Zaleski, S., 2013. Instability regimes in the primary breakup region of planar

AN US

coflowing sheets. Journal of Fluid Mechanics 736, 150–176.

Gepperth, S., Guildenbecher, D., Koch, R., Bauer, H.J., 2010. Pre-filming primary atomization: Experiments and modeling, in: ILASS Europe, 23rd Annual Conference on Liquid Atomization and Spray Systems, Brno, Czech Rep.

M

Gepperth, S., Koch, R., Bauer, H.J., 2013. Analysis and comparison of primary droplet characteristics in the near field of a prefilming airblast atomizer, in:

ED

Proceedings of ASME Turbo Expo 2013: Turbine Technical Conference and Exposition, San Antonio, Texas, USA.

PT

Gepperth, S., Müller, A., Koch, R., Bauer, H.J., 2012. Ligament and droplet characteristics in prefilming airblast atomization, in: ICLASS, 12th Triennial International Annual Conference on Liquid Atomization and Spray Systems,

CE

Heidelberg, Germany.

AC

Grosshans, H., Fuchs, L., 2011. Development of a combined vof-lpt method to simulated two-phase flows in various regimes, in: TSFP-7, Ottawa, Canada.

Hashmi, A.A., Dullenkopf, K., Koch, R., Bauer, H.J., 2010. Cfd methods for shear driven liquid wall films, in: Proceedings of ASME Turbo Expo 2010: Power for Land, Sea and Air, Glasgow, UK. pp. 1283–1291.

48

ACCEPTED MANUSCRIPT

Herbert, D.A., Schmidt, D.P., Knaus, D.A., Phililips, S., Magaria, P.J., 2008. Parallel vof spray droplet identification in an unstructured grid, in: ILASS Americas, 21st Annual Conference on Liquid Atomization and Spray Systems,

CR IP T

Orlando, FL, USA. Hoefler, C., Braun, S., Koch, R., Bauer, H.J., 2013. Modeling spray formation in gas turbines - a new meshles approach. Journal of Engineering for Gas Turbines and Power 135.

Iwamoto, K., 2002. Database of fully developed turbulent channel flow, Univer-

AN US

sity of Tokyo, Department of Mechanical Engineering, Tokyo.

Jasuja, A.K., 1981. Airblast atomization of alternative liquid petroleum fuels under high pressure conditions. Journal of Engineering for Power 103, 514– 518.

Jasuja, A.K., 2006. Behaviour of aero-engine airblast sprays in practical envi-

M

ronments, in: ICLASS, 10th Triennial International Annual Conference on Liquid Atomization and Spray Systems, Kyoto, Japan.

ED

Kapulla, R., Tuchtenhagen, J., Müller, A., Dullenkopf, K., Bauer, H.J., 2008. Droplet sizing performance of different shadow sizing codes, in: GALA Fach-

PT

tagung Lasermethoden in der Strömungsmesstechnik. Kim, K.S., Kim, S.S., 1994. Drop sizing and depth-of-field correction in tv

CE

images. Atomization and Sprays 4, 65–78. Kline, S.J., McClintock, F.A., 1953. Describing uncetainties in single-sample experiments. Mechanical Engineering 75, 3–8.

AC

Lee, S.Y., Kim, Y.D., 2004. Sizing of spray particles using image processing technique. KSME International Journal 18, 879–894.

Lefebvre, A.H., 1980. Airblast atomization. Progress in Energy and Combustion Sience 6, 233–261.

49

ACCEPTED MANUSCRIPT

Lefebvre, A.H., Miller, D., 1966. The deveoplment of an air blast atomizer for gas turbine application. Technical Report CoA. Report AERO No. 193. The college of aeronautics cranfield.

CR IP T

Ling, Y., Zaleski, S., 2014. Numerical investigation of spray formation and evolution in gas-assisted atomization, in: ILASS Europe, 26th Annual Conference on Liquid Atomization and Spray Systems, Bremen, Germany.

Marmottant, P., Villermaux, E., 2004. On spray formation. Journal of Fluid Mechanics 498, 73–111.

AN US

Moser, R.D., Kim, J., Mansour, N.N., 1999. Direct numerical simulation of turbulent channel flow up to Reτ = 590. Physics of Fluids 11, 943–945.

Müller, A., 2015. Experimentelle Untersuchung des Zerstäubungsverhaltens luftgestützter Brennstoffdüsen bei oszillierenden Strömungen. Ph.D. thesis. Technologie (KIT).

M

Institut für Thermische Strömungsmaschinen (ITS), Karlsruher Institut für

Müller, A., Hehle, M., Koch, R., Bauer, H.J., 2005. Zerstäubungsverhalten von

ED

airblastdüsen bei oszillierenden strömungen (performance of airblast atomisers in unsteady flow conditions). VDI-Bericht, 22. Deutscher Flammentag:

PT

Verbrennung und Feuerung 1888, 249–257. Müller, A., Koch, R., Bauer, H.J., Hehle, M., Schäfer, O., 2006. Performance of prefilming airblast atomizers in unsteady flow conditions, in: Proceedings

CE

of ASME Turbo Expo 2006: Power for Land, Sea and Air, Barcelona, Spain.

AC

Müller, A., Schumann, F., Dullenkopf, K., Bauer, H.J., 2007. Analysis of droplet wall interaction using advanced image processing techniques, in: ILASS Europe, 21th Annual Conference on Liquid Atomization and Spray Systems, Mugla, Turkey.

NSaTC, 2010. National aeronautics research and development plan. Technical Report. Executive Office of the President of the United States of America.

50

ACCEPTED MANUSCRIPT

Olsson, E., Kreiss, G., 2005. A conservative level set method for two phase flow. Journal of Computational Physics 210, 225–246. Olsson, E., Kreiss, G., Zahedi, S., 2007. A conservative level set method for two

CR IP T

phase flow ii. Journal of Computational Physics 225, 785–807.

Pope, S.B., 2000. Turbulent Flows. Cambridge University Press, Cambridge.

Pringuey, T., Cant, R.S., 2014. Robust conservative level set method for 3d

mixed-element meshes - application to les of primary liquid-sheet breakup.

AN US

Communications in Computational Physics 16, 403–439.

Rizk, N.K., Lefebvre, A.H., 1983. Influence of atomizer design features on mean drop size. AIAA Journal 21, 1139–1142.

Rizkalla, A.A., Lefebvre, A.H., 1975. The influence of air and liquid properties on airblast atomization. Journal of Fluids Engineering 97, 316–320.

M

Rusche, H., 2002. Computational Fluid Dynamics of Dispersed Two-Phase Flows at High Phase Fractions. Ph.D. thesis. University of London.

ED

Sander, W., Weigand, B., 2008. Direct numerical simulation and analysis of instability enhancing parameters in liquid sheets at moderate reynolds numbers.

PT

Physics of Fluids 20, 053301.

Sattelmayer, T., Wittig, S., 1986. Internal flow effects in prefilming airblast atomizers: mechanisms of atomization and droplet spectra. Journal of Engi-

CE

neering for Gas Turbines and Power 108, 465–472.

AC

Sauer, B., 2014. Direct numerical simulation of the primary breakup of aircraft engine related two-phase flows. Ph.D. thesis. Technische Universität Darmstadt. Darmstadt, Germany.

Sauer, B., Sadiki, A., Janicka, J., 2016. Embedded dns concept for simulating the primary breakup of an airblast atomizer. Atomization and Sprays 26, 187–215.

51

ACCEPTED MANUSCRIPT

Spyrou, N., 2011. Detaillierte numerische Simulationen von flugtriebwerksrelevanten Zweiphasenströmungen. Ph.D. thesis. Technische Universität Darmstadt.

Fluid Mechanics 43, 399–426.

CR IP T

Tropea, C., 2011. Optical particle characterization in flows. Annual Review of

Tropea, C., Damaschke, N., 2000. New developments in phase doppler anemometry. Journal of Visualization 2, 273–279.

Villedieu, P., Blanchard, G., Zuzio, D., 2013. Numerical simulation of primary

AN US

atomization of a sheared liquid sheet. part2: Comparison with experimental results, in: ILASS Europe, 25th European Conference on Liquid Atomization and Spray Systems, Chania, Greece.

Volz, M., Konle, M., Gebretsadik, M., Habisreuther, P., Zarzalis, N., 2015. Investigation of a prefilming airblast atomizer with respect to surface stripping,

M

in: ASME Turbo Expo 2015: Turbine Technical Conference and Exposition, Montreal, Quebec, Canada.

ED

Wilfert, G., Sieber, J., Rolt, A., Baker, N., Touyeras, A., Colantuoni, S., 2007. New environmental friendly aero engine core concepts, in: ISABE-2007-1120.

PT

Yoshizawa, A., Horiuti, K., 1985. A statistically-derived subgrid-scale kinetic energy model for the large-eddy simulation of turbulent flows. Journal of the

CE

Physical Society of Japan 54, 2834–2839. Zuzio, D., Estivalezes, J.L., Villedieu, P., Blanchard, G., 2013. Numerical sim-

AC

ulation of primary and secondary atomization. Comptes Rendus Mecanique 341, 15–25.

52

ACCEPTED MANUSCRIPT

List of Figures Experimental setup (all units are represented in mm) . . . . . . .

8

2

Liquid accumulation at the atomizing edge . . . . . . . . . . . .

9

3

Detected liquid accumulation and droplets after data processing .

10

4

Calibrated droplet diameter as a function of the out-of-focus po-

CR IP T

1

sition and reference diameter . . . . . . . . . . . . . . . . . . . . 5

12

Calibrated normalised intensity gradients as a function of the

out-of-focus position and reference diameter . . . . . . . . . . . . 6

15

Calculation of the real droplet diameter based on the normalized

16

7

The embedded DNS domain for an annular airblast atomizer . .

17

8

Geometry simplification of the airblast atomizer . . . . . . . . . .

18

9

Application of the eDNS concept to a planar airblast atomizer .

18

10

Impact of the inlet boundary condition variation on the pri-

AN US

intensity gradient and the measured droplet diameter . . . . . . .

M

mary breakup. Gas phase: u = 50 m/s, ρ = 3.9 kg/m3 , ν = 4.7 × 10−6 m2 /s. Liquid phase: u = 0.5 m/s, ρ = 650 kg/m3 ,

19

Geometry and grid of the turbulent channel . . . . . . . . . . . .

21

12

Computational domain of the planar prefilmer . . . . . . . . . . .

24

13

Reducing three-dimensional DNS data to a plot of the interface .

27

14

Experimental field of view and the used image section of the size

PT

ED

.

11

ν = 2.9 × 10−6 m2 /s, σ = 0.0085 kg/s2 . Prefilmer: h = 50 µm.

of the DNS domain for the comparison of the breakup phenomena 30 Disintegration sequence at the atomizing edge with ligament for-

CE

15

mation. From left to right: Sideview experiment, topview exper-

AC

iment, topview simulation, sideview simulation . . . . . . . . . .

16

33

Disintegration sequence at the atomizing edge with the formation

of bag like structures. From left to right: Sideview experiment, topview experiment, topview simulation, sideview simulation . . 17

34

Peak detection for the phase interface (a) and the effect of a minimum distance of 50 µm (b) for the analysed operating point

53

36

ACCEPTED MANUSCRIPT

18

Time displacement of the phase interface and ambiguities by its 37

19

Distribution of the ligament length . . . . . . . . . . . . . . . . .

39

20

Distribution of the deformation velocity . . . . . . . . . . . . . .

39

21

Sensitivity of breakup length and deformation velocity dependent

CR IP T

determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

on the liquid fraction α . . . . . . . . . . . . . . . . . . . . . . . 22

Representation of bubbles in the experiment (a) and the simula-

tion with α=0.5 (b) or α=0.05 (c) . . . . . . . . . . . . . . . . . 23

AN US

43

Droplet distribution for diameter and streamwise velocity: Comparison between experiment and simulation . . . . . . . . . . . .

26

42

Small structure properties extracted at the eDNS outflow: droplet mass and number of droplets . . . . . . . . . . . . . . . . . . . .

25

40

Droplet size distribution: Comparison between experiment and

simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

40

43

Scatter plot for the droplet diameter dd and the streamwise droplet

AC

CE

PT

ED

M

velocity uz . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

54

44

ACCEPTED MANUSCRIPT

List of Tables 1

Characteristic parameters of the optical configuration . . . . . .

2

Non-dimensional grid spacings for the turbulent channel. Reg =

13 22

3

Operating conditions . . . . . . . . . . . . . . . . . . . . . . . . .

29

4

Characteristic values for the accumulated liquid . . . . . . . . . .

38

AC

CE

PT

ED

M

AN US

CR IP T

13, 333, Reτ = 777. . . . . . . . . . . . . . . . . . . . . . . . . . .

55