19
The Influence of Density Ratio on the Primary Atomization of a Turbulent Liquid Jet in Crossflow M. Herrmann School of Mechanical, Aerospace, Chemical, and Materials Engineering Arizona State University, Tempe, AZ 85287-6106, USA. Abstract In this paper we study the impact of density ratio on turbulent liquid jet in crossflow penetration and atomization if all other characteristic parameters, i.e. momentum flux ratio, jet and crossflow Weber and Reynolds numbers, are maintained constant. We perform detailed simulations of the primary atomization region using the refined level set grid method to track the motion of the liquid/gas phase interface. We employ a balanced force, interface projected curvature method to ensure high accuracy of the surface tension forces, use a multi-scale approach to transfer broken off, small scale nearly spherical drops into a Lagrangian point particle description allowing for full two-way coupling and continued secondary atomization, and employ a dynamic Smagorinsky large eddy simulation approach in the single phase regions of the flow to describe turbulence. We compare simulation results obtained previously using a liquid to gas density ratio of 10 for a momentum flux ratio 6.6, Weber number 330, and Reynolds number 14000 liquid jet injected into a Reynolds number 740,000 gaseous crossflow to those at a density ratio of 100, a value typical for gas turbine combustors. The results show that the increase in density ratio results in a noticeable increase in jet penetration, change in drop size distribution resulting from primary atomization, change in drop velocities generated by primary atomization in the crossflow and jet direction, but virtually no change in drop velocities in the transverse direction. Keywords: atomization, level set, jet in crossflow, jet penetration, drop sizes 1. Introduction The atomization of turbulent liquid jets injected into fast moving, subsonic gaseous crossflows is an important application for example in gas turbines, ramjets, and augmentors. It is a highly complex process, that has been extensively studied experimentally over the past decades. Early studies of the atomization of non-turbulent liquid jets in crossflows have recently been reviewed in [1], whereas newer studies of this case include the work reported in Preprint submitted to Elsevier May 27, 2010

The In uence of Density Ratio on the Primary Atomization of a Turbulent Liquid Jet …multiphase.asu.edu/paper/cs_2010.pdf · 2010. 8. 9. · Besides jet penetration, a change in

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • The Influence of Density Ratio on the Primary Atomization of a Turbulent LiquidJet in Crossflow

    M. Herrmann

    School of Mechanical, Aerospace, Chemical, and Materials EngineeringArizona State University, Tempe, AZ 85287-6106, USA.

    Abstract

    In this paper we study the impact of density ratio on turbulent liquid jet in crossflow penetration and atomization

    if all other characteristic parameters, i.e. momentum flux ratio, jet and crossflow Weber and Reynolds numbers, are

    maintained constant. We perform detailed simulations of the primary atomization region using the refined level set

    grid method to track the motion of the liquid/gas phase interface. We employ a balanced force, interface projected

    curvature method to ensure high accuracy of the surface tension forces, use a multi-scale approach to transfer broken

    off, small scale nearly spherical drops into a Lagrangian point particle description allowing for full two-way coupling

    and continued secondary atomization, and employ a dynamic Smagorinsky large eddy simulation approach in the

    single phase regions of the flow to describe turbulence. We compare simulation results obtained previously using a

    liquid to gas density ratio of 10 for a momentum flux ratio 6.6, Weber number 330, and Reynolds number 14000

    liquid jet injected into a Reynolds number 740,000 gaseous crossflow to those at a density ratio of 100, a value typical

    for gas turbine combustors. The results show that the increase in density ratio results in a noticeable increase in jet

    penetration, change in drop size distribution resulting from primary atomization, change in drop velocities generated

    by primary atomization in the crossflow and jet direction, but virtually no change in drop velocities in the transverse

    direction.

    Keywords: atomization, level set, jet in crossflow, jet penetration, drop sizes

    1. Introduction

    The atomization of turbulent liquid jets injected into fast moving, subsonic gaseous crossflows is an important

    application for example in gas turbines, ramjets, and augmentors. It is a highly complex process, that has been

    extensively studied experimentally over the past decades. Early studies of the atomization of non-turbulent liquid

    jets in crossflows have recently been reviewed in [1], whereas newer studies of this case include the work reported in

    Preprint submitted to Elsevier May 27, 2010

  • [2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12]. Most experimental work has focused on jet penetration, including both the column

    trajectory and the resulting spray penetration under ambient, atmospheric conditions. Some recent jet penetration

    correlations under these conditions can be found in [3, 10, 11, 12]. Under non-atmospheric conditions, jet penetration

    correlations have been derived in [2, 13]. The identified characteristic parameters determining jet penetration from

    these studies are the momentum flux ratio [12], momentum flux ratio and crossflow Weber number [11], momentum

    flux ratio and water to injected liquid viscosity ratio [3], momentum flux ratio, crossflow Weber number, and water to

    injected liquid viscosity ratio [10], momentum flux ratio, crossflow Weber and Reynolds numbers [2], and momentum

    flux ratio, crossflow Weber number, and channel pressure [13]. Since gaseous density is proportional to channel

    pressure, only the latter correlation contains the impact of density ratio on liquid jet penetration directly. However,

    although momentum flux ratio and crossflow and jet Weber numbers were approximately held constant, at least one

    of the Reynolds numbers could not have been maintained, thus potentially overlapping two separate effects. It should

    be pointed out that some other studies exist that vary the channel pressure from its atmospheric value, however,

    these studies consider modified configurations, for example airblast assisted atomization [14], or did not result in

    modifications to the liquid jet core penetration correlations [15].

    All the above studies at non-atmospheric conditions use as a starting point correlations obtained at atmospheric

    conditions, since there is an abundance of data there. Density ratios found in applications like gas turbine combustors

    are thus approached from above, i.e. atmospheric conditions result in too large density ratios and non-atmospheric

    experiments analyze the effect of increasing the crossflow channel pressure and thereby reducing the density ratio in

    steps to those values representative of gas turbine combustors [13].

    Detailed numerical simulations of turbulent atomization processes, on the other hand, are more easily performed

    at density ratios smaller than those found in gas turbine combustors due to often times encountered numerical

    instabilities associated with the density discontinuity on highly deformed interfaces. Yet how representative are

    these lower density ratio simulations of the actual application in gas turbine combustors? We aim to address this

    question by comparing detailed simulation results of the primary atomization region at two different density ratios,

    one lower, the other representative of gas turbine injectors, under the constraint that all characteristic numbers, i.e.

    momentum flux ratio, crossflow and jet Weber numbers, and crossflow and jet Reynolds numbers are kept constant.

    This effectively isolates the impact of density ratio, a feature not yet reproduced experimentally.

    2

  • Besides jet penetration, a change in density ratio, while maintaining all other characteristic numbers, might also

    impact the primary atomization mechanism of the jet, and thus the resulting primary atomization drop size and

    velocity distributions.

    The outline of this paper is the following: after summarizing the governing equations, the numerical methods

    employed to solve them are briefly outlined. Then, simulation results are presented and the impact of density ratio

    on the atomization process is discussed.

    2. Governing equations

    The equations governing the motion of an unsteady, incompressible, immiscible, two-fluid system are the Navier-

    Stokes equations,

    ∂u

    ∂t+ u · ∇u = −1

    ρ∇p+ 1

    ρ∇ ·

    (µ(∇u +∇Tu

    ))+

    1

    ρT σ , (1)

    where u is the velocity, ρ the density, p the pressure, µ the dynamic viscosity, and T σ the surface tension force

    which is non-zero only at the location of the phase interface xf . Furthermore, the continuity equation results in a

    divergence-free constraint on the velocity field, ∇ · u = 0. The phase interface location xf between the two fluids is

    described by a level set scalar G, with G(xf , t) = 0 at the interface, G(x, t) > 0 in fluid 1, and G(x, t) < 0 in fluid

    2. Differentiating this with respect to time yields,

    ∂G

    ∂t+ u · ∇G = 0 . (2)

    Assuming material properties to be constant within each fluid, they can be calculated at any point x from

    α(x) = H(G)α1 + (1−H(G))α2 (3)

    where α is either density or viscosity, indices 1 and 2 denote values in fluid 1, respectively 2, and H is the Heaviside

    function. The surface tension force T σ can be expressed as

    T σ(x) = σκδ(x− xf )n = σκδ(G)|∇G|n , (4)

    with σ the surface tension coefficient and interface normal n and curvature κ directly calculated from the level set

    scalar.

    3

  • 3. Numerical methods

    To keep track of the position and motion of the phase interface, we solve all level set related equations using the

    Refined Level Set Grid (RLSG) on a separate, equidistant Cartesian grid using a dual-narrow band methodology for

    efficiency. This so-called G-grid is overlaid onto the flow solver grid on which the Navier-Stokes equations are solved

    and can be independently refined, providing high resolution of the tracked phase interface geometry. Details of the

    resulting method and extensive verification results are reported in [16].

    The flow solver used to solve the incompressible two-phase Navier-Stokes equations on unstructured grids using

    a finite volume balanced force algorithm is Cascade Technologies’ CDP. In the single phase regions, the employed

    scheme conserves the kinetic energy discretely and turbulence is modeled using a dynamic Smagorinsky LES model.

    However, none of the terms arising from filtering the phase interface, like subfilter surface tension or subfilter liquid

    volume fraction transport, are modeled. The approach instead relies on resolving all relevant scales at the phase

    interface, thus ideally reverting to a DNS at the phase interface. It is for this reason that the highest grid resolutions

    in the simulations reported here are provided in the vicinity of the phase interface. As such, the current simulation

    results are a combination of LES in the single phase regions and an under-resolved DNS on the two-dimensional

    subset of the phase interface [17]. The flow solver CDP and the RLSG interface tracking software are coupled using

    the parallel multi-code coupling library CHIMPS [16, 18].

    Finally, resolving the entire phase interface geometry by tracking the phase interface associated with each at-

    omized drop quickly becomes prohibitively expensive. Instead, we follow a multi-scale coupled Eulerian/Lagrangian

    procedure in that we track the phase interface by the Eulerian level set method in the near injector primary atomiza-

    tion region and transfer broken-off, nearly spherical liquid structures into a Lagrangian point particle description [19].

    To this end, the level set solution field is searched at the end of every time step to identify all unique G > 0 volumes.

    If the volume is smaller than a given threshold value, it’s eccentricity measure defined as the maximum distance

    from the structure’s surface to its center of mass scaled by the radius of the equivalent volume sphere is calculated.

    If the eccentricity measure is smaller than a given threshold value, the liquid structure is removed from the level set

    representation, by simply switching the sign of the level set values within the structure, thereby resetting the volume

    fraction and densities of the structure’s occupied volume to the gaseous values. A Lagrangian particle is added to

    the flow solver at the structure’s center of mass with the volume averaged velocity, thereby preserving the removed

    4

  • structure’s mass and momentum. In the Lagrangian description, full two-way momentum coupling between the drop

    and continuous phase is used [20], including a stochastic secondary atomization model [21]. However, the cell volume

    occupied by the Lagrangian drops is not explicitly taken into account and neither drop/drop nor drop/tracked phase

    interface collisions are modeled. But, as long as liquid structures have not been transferred into the Lagrangian

    description, i.e. they are still tracked by the level set scalar, cell volume effects and all collisions are fully captured.

    4. Computational Domain and Operating Conditions

    The case analyzed in this paper is based on one studied experimentally by Brown & McDonell [4]. Table 1

    summarizes the operating conditions and resulting characteristic numbers. Results obtained for the density ratio 10

    case were reported previously in [17].

    The computational domain (−25D . . . 50D × 0 . . . 25D × −10D . . . 10D) is smaller than the channel used in the

    experiment (−77D . . . 127D × 0 . . . 54D ×−27D . . . 27D). However, simulations using the full experimental channel

    geometry were conducted to verify that the reduced computational domain does not impact the reported results.

    Furthermore, the details of the injector geometry are taken into account by performing at a constant momentum

    flux ratio a turbulent single phase LES of the full geometry and storing the resulting velocities in a database for use

    as the injector exit-plane inflow boundary condition in the atomization simulations [17].

    We previously reported on a grid refinement study for the ρj/ρc = 10 case using three different grid resolutions

    [17]. The obtained results indicate that jet penetration can be well predicted even on the coarsest grid, column and

    surface breakup modes can be observed on all three grids, and that grid independent drops resulting from primary

    atomization can be obtained for drops resolved by at least 2 grid nodes of the RLSG grid. Unless otherwise indicated,

    all results in the following are obtained using the medium resolution grid c12 of [17] as he base grid. It resolves the

    injector diameter by 32 control volumes using a static mesh refinement strategy in the injector region in the flow

    solver and 64 grid nodes in the RLSG solver. The resulting mesh sizes are 21 million control volumes for the flow

    solver and 13 million active RLSG grid nodes. At t = 0, the liquid jet is initialized in the computational domain by

    a small cylindrical section of length D capped by a half-sphere, protruding into the crossflow channel. The threshold

    volume used to transfer drops from the level set to the Lagrangian description is 0.14µl and the chosen threshold

    eccentricity measure is 1.5.

    5

  • In the following, unless otherwise indicated, results are reported in dimensionless quantities, with uj and D being

    the reference velocity and length. Thus, non-dimensional coordinates are the same in both density cases, whereas

    non-dimensional times and velocities are different by a factor 3.16. After an initial startup phase, the atomizing jets

    reach a statistical steady state of the level set tracked liquid mass at around t = 20 in the low density ratio case and

    t = 10 in the high density ratio case. All results reported in the following are obtained only after this initial startup

    phase.

    5. Results and Discussion

    Figure 1 shows instantaneous snapshots of the surface geometry of the two cases. Jet penetration appears different,

    with the higher density case being less bent in the crossflow direction, and the liquid core not extending as far in

    that direction.

    To allow for a detailed quantitative analysis of the mean jet penetration, the following procedure is adopted to

    extract such data from time snapshots of the level set solution field. Since the interest here is on time averaged data,

    it should be pointed out that a simple time averaging of the level set scalar is doomed to fail and give inaccurate

    results, since such an approach violates the inherent symmetries of the level set equation [22]. Instead, we base

    our averaging tool on the spatial filters for level set scalars proposed in [23], applied to temporal filtering instead of

    spatial filtering. The idea in this approach is to perform the averaging on a Heaviside transform of the level set scalar

    instead of on the level set scalar itself. This, in essence, results in a scalar that describes the threshold probability

    of finding liquid at a given location (the probability is x% or greater). Additionally, this scalar is consistent with

    a mean liquid volume fraction. We define the 50% probability iso-surface to be the location of the mean surface,

    although, in principle, any other value could be used.

    Figure 2 presents the position of the liquid core’s mean center of mass and windward and lee-side mean interface

    position in the injector’s center plane. The center of mass position was determined by an iterative procedure evalu-

    ating the center of mass in planes normal to its trajectory. Also shown are power law fits, where y/D = 3.3(x/D)0.33

    and y/D = 3.8(x/D)0.40 fit the center of mass trajectory in the low and high density ratio case, respectively. The

    50% probability leading edge can be fitted by y/D = 3.18(x/D)0.49, respectively y/D = 3.4(x/D)0.45. The power

    law fits obtained from experiments are y/D = 3.5(x/D)0.50 [12] and y/D = 3.6(x/D)0.39 [24].

    6

  • It should, be pointed out, that the pre-factor and exponent in the power-law fit are very sensitive to the chosen

    value of the probability iso-line. Choosing values ranging from 10% up to 90% results in power law fit pre-factors

    ranging from 2.74 to 3.50 and exponents from 0.39 to 0.52 in the low density ratio case, and 3.1 to 3.7 with exponents

    from 0.36 to 0.48 in the high density ratio case. This significant spread in parameters highlights a potential weakness

    in the way windward edge trajectories, and therefore, jet penetration is experimentally measured. The results indicate

    that the power law parameters are a strong function of the threshold value used to identify the windward edge. A

    more objective measure for jet penetration would be the jet’s center of mass, which is, however, hard to measure

    experimentally. Nonetheless the obtained leading edge correlations are within the range determined experimentally,

    thus providing some level of validation of the simulation data. To verify the numerical solution, Fig. ?? shows the

    result of a grid dependency study using a factor 2 coarser grid in each direction compared to the base grid. Windward

    and lee side mean interface position show practically no change between the two grids, thus verifying the presented

    solutions.

    Figure 3 shows the discrete number and volume frequency distributions of drops generated by primary atomization

    in the two density ratio cases. Since there is no precise definition of the demarcation between primary and secondary

    atomization, we define primary atomization as all topology change events resulting in liquid structures smaller than

    a threshold volume and nearly spherical in shape as defined by the eccentricity measure previously stated, i.e. drops

    triggering a transfer from the level set tracked representatioan to the Lagrangian description. The threshold volume

    used in the simulations corresponds to a maximum spherical drop size of d = 1/2, thereby allowing for the transfer of

    very large structures. The used eccentricity threshold however prevents the transfer of ligament or sheet structures

    even if they are small in volume and broken off from the main liquid column. Instead, the drops resulting from the

    breakup of these highly non-spherical structures are mostly spherical and thus are considered to be the result of

    primary atomization since they typically trigger a transfer.

    The discrete frequency distribution is obtained by binning the drop sizes using 20 bins of equal logarithmic drop

    size between the smallest drop size resolved by at least 2 G-grid points and d = 1/4 which is larger than the largest

    transferred drop throughout either simulation. The distribution is normalized by the total number of drops in this

    size range, 24,745 drops in the low density ratio case and 22,233 drops in the high density ratio case.

    It should be pointed out that it is practically impossible to obtain a fully grid independent number frequency

    7

  • distribution using fixed grid interface capturing/tracking schemes like the Volume-of-Fluid or the level set method.

    This is due to the fact that in these methods, topology change events are automatically initiated if two opposing

    front segments enter into the same computational grid cell. This automatic topology change event is always grid

    dependent and the local mesh size acts as an inherent topology change parameter, unless a dedicated topology

    change model based on micro-scale physics like Van der Waals forces would be employed [25]. Although the exact

    moment of topology change might not be predicted accurately without such a micro-scale model, nevertheless the

    error in volume of the broken-off liquid structure is at most of the order of the local mesh volume and thus small

    compared to drops resolved by several grid points pre-transfer. Indeed grid independence of such larger drops has

    been demonstrated in the past, see [17, 26] for examples. We thus consider only drops resolved by at least 2 G-grid

    points in all of the presented results. Finally, to ensure that results of the two density ratios are comparable, we

    necessarily have to use the same inherent topology change model in each case, i.e. the same grid resolution, as is the

    case in the present study.

    The discrete frequency distributions show that changing the density ratio does have a noticeable impact on

    drop sizes resulting from primary atomization. Increasing the density ratio decreases the number of larger drops

    significantly and yields a somewhat narrower distribution. This observation is consistent with the Sauter mean and

    standard deviation calculated from the binned data giving d32 = 0.0762 and d32 = 0.0575 with σ = 8.21 · 10−3 and

    σ = 7.02 · 10−3 in the low respective high density ratio case. Note that these values are calculated using the discrete

    number frequency distribution in the given size range only, so only the trend should be considered, not the actual

    quantitative values.

    At first glance the observed trend that an increase in density ratio should result in a decrease in the mean drop size

    is somewhat puzzling. One would expect that an increase in density ratio should reduce the impact of the gas phase

    aerodynamic effect, thereby producing larger mean drops instead of smaller ones. This line of thinking however

    neglects the fact that the simulations were carried out maintaining as constant all other characteristic numbers,

    thereby isolating the effect of density ratio only. To keep for example the momentum flux ratio constant, we have to

    decrease the jet velocity if we increase the jet density to increase the density ratio, see Tab. 1. This will necessarily

    change the relative velocity between the jet and the crossflow in such a way that a rough estimate of the most

    unstable wave length of a Kelvin-Helmholtz type instability would in fact indicate that the wave length decreases

    8

  • when increasing the density ratio from 10 to 100, thus resulting in potentially smaller drops if one were to assume a

    Kelvin-Helmholtz type instability as the dominant instability mode leading to breakup of the liquid column [27].

    Figure 4 shows the location at which drops are generated by primary atomization, i.e. the location of transfer

    from the level set to the Lagrangian representation. In the high density ratio case, atomization occurs significantly

    closer to the injector, and the region of generation is significantly less bend in the crossflow direction. This can

    be explained by the different trajectories of the jets center of mass in the two density ratio cases. Theprimary

    atomization is also more compact in the transverse direction in the higher density ratio case. Although primary

    atomization seems to occur earlier, the transverse spread of the resulting spray is increased when the density ratio

    increases as can be seen in Fig. 1 showing top view snapshots of the developing spray. This trend is consistent with

    experimental observations showing an increase in spray area and plume width with increasing density ratios [2].

    Finally, Figs. 5-7 show the discrete number frequency distribution of primary atomization drop velocities at

    their moment of transfer from the level set to the Lagrangian representation. Data like this is of importance to

    modeling efforts of the primary atomization process that inject drops along projected mean liquid core paths with

    a given velocity [28]. The distributions have been generated by first binning drops into 6 size classes between

    2∆xG < d < 0.25 of equal logarithmic size named d0 to d5. Within each size class, the velocity data in each direction

    is binned into 20 equally sized bins between the minimum and maximum velocities in the respective directions. Total

    number of drops used in each size class for the low density ratio case are 15,594 for d0, 645 for d3, and 139 for d4,

    and for the high density case 14,225 for d0, 178 for d3, and 7 for d4.

    Figure 5 shows that density ratio has a significant impact on the crossflow direction drop velocity. Here scaling is

    performed with the gaseous crossflow velocity as the relevant velocity scale for this direction. In each case, velocities

    for large drops converge to a uniform velocity, consistent with experimental observations [12]. The value however is

    significantly different in the two cases. In the low density ratio case, the asymptotic velocity is close to 1, whereas in

    the high density ratio case the value is roughly 0.25. A similar trend in drop velocity has been reported by Elshamy

    [13], although measured significantly further downstream of the injector at x/D = 40. The observed difference is

    easily explained by the lower relative Stokes number of the liquid in the low density ratio case.

    Small drops, on the other hand exhibit a broad range of crossflow velocities, with a considerable number moving

    upstream at the moment of primary atomization. This is due to the fact that the liquid jet generates significant

    9

  • recirculation regions in its wake and small drops, due to their low Stokes number will follow the vortices there more

    easily resulting in potential upstream movement. This behavior is of significance, since residence times of drops

    caught in the recirculation region of the wake will be increased and combined with the fact that these drops tend to

    be small, they might provide a gaseous fuel source near the injector in combustion applications due to evaporation.

    Furthermore, the variance of the crossflow velocity distribution is noticeably larger in the low density ratio case,

    explainable again by the difference in Stokes numbers.

    Figure 6 shows on the top the jet direction velocities scaled with the liquid jet velocity as the seemingly relevant

    velocity scale in this direction. Density ratio appears to again have a significant impact with the high density

    ratio case showing a larger mean value, again due to the relative larger Stokes number, but also a much broader

    distribution. This might be due to the fact that the high density ratio case better retains the initial range of velocity

    fluctuations from the turbulent flow of the injector. Interestingly, however, if one scales the jet direction velocity

    with the cross flow velocity as shown in the bottom of Fig. 6, the two distributions nearly collapse. This indicates a

    scaling with the square root of the density ratio, since the momentum flux ratio is maintained constant. Furthermore,

    the distributions show a considerable number of drops moving in the opposite direction of the injected jet at their

    moment of primary atomization, again explainable by the turbulent wake region behind the injected jet.

    Finally Fig. 7 shows the transverse direction velocities scaled with the crossflow velocity as the relevant velocity

    scale. As expected, the distributions are symmetric, and appear not to be predominantly a function of density

    ratio. Larger drops seem to exhibit a more uniform velocity, however this trend is much less pronounced than in the

    crossflow direction.

    6. Conclusions

    Detailed simulation results of the primary atomization of a turbulent liquid jet injected into a subsonic gaseous

    crossflow have been presented, studying the impact of density ratio on the primary atomization region. Results were

    obtained using the same numerical topology change model given by the same local grid spacing in both cases. The

    simulation results indicate that the density ratio has a noticeable effect on primary atomization, even if all other

    relevant characteristic numbers are being kept constant. Mean jet penetration, location of primary atomization, drop

    sizes, crossflow and jet direction drop velocities at the moment of primary atomization are impacted by the change

    10

  • in density ratio. The simulation results, however, show that transverse direction drop velocities are not significantly

    changed.

    The obtained results can form a starting point for improving modeling attempts of jet in crossflow atomization

    based on the idea of injecting drops along projected mean liquid core paths with given velocity distributions [28].

    Acknowledgments

    This work was supported in part by CASCADE Technologies Inc. under NavAir SBIR N07-046. The author would

    like to thank M. Arienti and M. Soteriou from United Technologies Research Center for many helpful discussions

    and participation in the NavAir SBIR as well as S. Hajiloo and F. Ham from Cascade Technologies Inc.

    References

    [1] C. Aalburg, B. van Leer, G. M. Faeth, K. A. Sallam, Properties of nonturbulent round liquid jets in uniform

    gaseous cross flows, Atom. Sprays 15 (3) (2005) 271–294.

    [2] A. Bellofiore, A. Cavaliere, R. Ragucci, Air density effect on the atomization of liquid jets in crossflow, Combust.

    Sci. and Tech. 179 (1-2) (2007) 319–342.

    [3] M. Birouk, C. O. Iyogun, N. Popplewell, Role of viscosity on trajectory of liquid jets in a cross-airflow, Atom.

    Sprays 17 (3) (2007) 267–287.

    [4] C. T. Brown, V. G. McDonell, Near field behavior of a liquid jet in a crossflow, in: ILASS Americas, 19th

    Annual Conference on Liquid Atomization and Spray Systems, Toronto, Canada, 2006.

    [5] C. T. Brown, U. M. Mondragon, V. G. McDonell, Investigation of the effect of injector discharge coefficient

    on penetration of a plain liquid jet into a subsonic crossflow, in: ILASS Americas 20th Annual Conference on

    Liquid Atomization and Spray Systems, ILASS Americas, Chicago, IL, 2007.

    [6] R. P. Fuller, P.-K. Wu, K. A. Kirkendall, A. S. Nejad, Effects of injection angle on atomizationof liquid jets in

    transverse ai, AIAA J. 38 (1) (2000) 64–72.

    [7] K. Lee, C. Aalburg, F. J. Diez, G. M. Faeth, K. A. Sallam, Primary breakup of turbulent round liquid jets in

    uniform crossflows, AIAA J. 45 (8) (2007) 1907–1916.

    11

  • [8] J. Mazallon, Z. Dai, G. M. Faeth, Primary breakup of nonturbulent round liquid jets in gas crossflows, Atom.

    Sprays 9 (3) (1999) 291–311.

    [9] C. L. Ng, R. Sankarakrishnan, K. A. Sallam, Bag breakup of nonturbulent liquid jets in crossflow, International

    Journal of Multiphase Flow 34 (3) (2008) 241–259.

    [10] J. N. Stenzler, J. G. Lee, D. A. Santavicca, W. Lee, Penetration of liquid jets in a cross-flow, Atom. Sprays 16

    (2006) 887–906.

    [11] S. M. Thawley, U. M. Mondragon, C. T. Brown, V. G. McDonell, Evaluation of column breakpoint and trajec-

    tory for a plain liquid jet injected into a crossflow, in: Proceedings of the 21st Annual Conference on Liquid

    Atomization and Spray Systems, ILASS Americas, Orlando, FL, 2008, pp. 1–11.

    [12] P. K. Wu, K. A. Kirkendall, R. P. Fuller, A. S. Nejad, Breakup processes of liquid jets in subsonic crossflows, J.

    Propul. Power 13 (1) (1997) 64–73.

    [13] O. M. Elshamy, Experimental investigations of steady and dynamic behavior of transverse liquid jets, Ph.D.

    thesis, University of Cincinnati (2007).

    [14] M. Y. Leong, V. G. McDonell, G. S. Samuelsen, Effect of ambient pressure on an airblast spray injected into a

    crossflow, J. Propul. Power 17 (5) (2001) 1076–1084.

    [15] J. Becker, C. Hassa, Breakup and atomization of a kerosene jet in crossflow at elevated pressure, Atom. Sprays

    11 (2002) 49–67.

    [16] M. Herrmann, A balanced force refined level set grid method for two-phase flows on unstructured flow solver

    grids, J. Comput. Phys. 227 (4) (2008) 2674–2706.

    [17] M. Herrmann, Detailed numerical simulations of the primary atomization of a turbulent liquid jet in crossflow,

    Journal of Engineering for Gas Turbines and Power 132 (6) (2010) 061506–10.

    URL http://link.aip.org/link/?GTP/132/061506/1

    [18] J. J. Alonso, S. Hahn, F. Ham, M. Herrmann, G. Iaccarino., G. Kalitzin, P. LeGresley, K. Mattsson, G. Medic,

    P. Moin, H. Pitsch, J. Schluter, M. Svard, E. V. der Weide, D. You, X. Wu, CHIMPS: A high-performance

    12

  • scalable module for multi-physics simulation, in: 42nd AIAA/ASME/SAE/ASEE Joint Propulsion Conference

    & Exhibit, no. 2006-5274 in AIAA-Paper, 2006.

    [19] M. Herrmann, A parallel eulerian interface tracking/lagrangian point particle multi-scale coupling procedure, J.

    Comput. Phys. 229 (2010) 745–759.

    [20] P. Moin, S. V. Apte, Large-eddy simulation of realistic gas turbine combustors, AIAA J. 44 (4) (2006) 698–708.

    [21] S. V. Apte, M. Gorokhovski, P. Moin, Les of atomizing spray with stochastic modeling of secondary breakup,

    Int. J. Multiphase Flow 29 (9) (2003) 1503–1522.

    [22] M. Oberlack, H. Wenzel, N. Peters, On symmetries and averaging of the G-equation for premixed combustion,

    Combust. Theory Modelling 5 (2001) 363–383.

    [23] M. Herrmann, DNS of turbulent primary atomization using a level set/vortex sheet method, in: ILASS Americas

    18th Annual Conference on Liquid Atomization and Spray Systems,Irvine, CA, 2005.

    [24] J. N. Stenzler, J. G. Lee, D. A. Santavicca, Penetration of liquid jets in a cross-flow, in: 41st Aerospace Sciences

    Meeting and Exhibit, no. AIAA 2003-1327 in AIAA-Paper, 2003.

    [25] M. Gorokhovski, M. Herrmann, Modeling primary atomization, Annual Review of Fluid Mechanics 40 (1) (2008)

    343–366. arXiv:http://arjournals.annualreviews.org/doi/pdf/10.1146/annurev.fluid.40.111406.102200.

    [26] M. Herrmann, On simulating primary atomization using the refined level set grid method, in: ILASS Americas

    21st Annual Conference on Liquid Atomization and Spray Systems, Orlando, FL, 2008.

    [27] M. Arienti, M. C. Soteriou, Time-resolved proper orthogonoal decomposition of liquid jet dynamics, Phys. Fluids

    21 (112104) (2009) 1–15.

    [28] M. Behzad, A. Mashayek, N. Ashgriz, A KIVA-based model for liquid jet in cross flow, in: ILASS Americas,

    22nd Annual Conference on Liquid Atomization and Spray Systems, no. ILASS2010-112, 2010.

    13

  • sim. 1 sim. 2

    density ratio R = ρj/ρc 10 100

    jet exit diameter D [mm] 1.3 1.3

    crossflow density ρc [kg/m3] 1.225 1.225

    jet density ρj [kg/m3] 12.25 122.25

    crossflow velocity uc [m/s] 120.4 120.4

    jet velocity uj [m/s] 97.84 30.94

    crossflow viscosity µc [kg/ms] 1.82 e-5 1.82e-5

    jet viscosity µj [kg/ms] 1.11e-4 3.5e-4

    surface tension coeff. σ [N/m] 0.07 0.07

    momentum flux ratio q 6.6 6.6

    crossflow Weber number Wec 330 330

    jet Weber number Wej 2178 2178

    crossflow Reynolds number Rec 5.7e5 5.7e5

    jet Reynolds number Rej 14079 14079

    Table 1: Operating conditions and characteristic numbers.

    14

  • Figure 1: Side (top) and top (bottom) view of jet in crossflow atomization at t=30, ρj/ρc = 10 (left) and 100 (right).

    y

    0

    1

    2

    3

    4

    -1 0 1 2 3

    y

    0

    1

    2

    3

    4

    -1 0 1 2 3

    y

    0

    1

    2

    3

    4

    -1 0 1 2 3

    y

    x0

    1

    2

    3

    4

    -1 0 1 2 3x

    Figure 2: Top: center of mass, windward and lee side 50% probability isoline (lines) and power law fits (symbols) in the injector midplane.

    Bottom: Impact of grid resolution: base grid (lines), coarse grid (symbols). ρj/ρc = 10 (left) and 100 (right).

    N/Nsum

    10-5

    0.0001

    0.001

    0.01

    0.1

    1

    0.03 0.05 0.1 0.2 0.3d

    10-4

    D3 N/Nsum

    10-7

    10-6

    10-5

    0.03 0.05 0.1 0.2 0.3d

    Figure 3: Discrete number (left) and volume (right) frequency distributions of primary atomization drops, ρj/ρc = 10 (circles, black)

    and 100 (squares, gray).

    15

  • Figure 4: Location of drop generation side view (center), top view (right), ρj/ρc = 10 (circles, black) and 100 (squares, gray).

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    -1 -0.5 0 0.5 1 1.5up /uc up/uc

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    -1 -0.5 0 0.5 1 1.5

    N/Nsum

    Figure 5: Primary atomization crossflow direction drop velocities for drop size classes d0 (left) and d4 (right) for ρj/ρc = 10 (black) and

    100 (gray).

    N/Nsum

    N/Nsum

    0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    0.35

    -1.5 -1 -0.5 0 0.5 1 1.5 2 2.50

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    0.35

    -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5vp /ujvp /uj

    vp /ucvp /uc

    0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    -0.5 0 0.5 10

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    -0.5 0 0.5 1

    Figure 6: Primary atomization jet direction drop velocities scaled by jet (top) and crossflow velocity (bottom) for drop size classes d0

    (left) and d3 (right) for ρj/ρc = 10 (black) and 100 (gray).

    16

  • N/Nsum

    0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    -1 -0.5 0 0.5 10

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    -1 -0.5 0 0.5 1wp/ucwp/uc

    Figure 7: Primary atomization transverse direction drop velocities for drop size classes d0 (left) and d3 (right) for ρj/ρc = 10 (black)

    and 100 (gray).

    17

  • List of Tables

    1 Operating conditions and characteristic numbers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

    18

  • List of Figures

    1 Side (top) and top (bottom) view of jet in crossflow atomization at t=30, ρj/ρc = 10 (left) and 100

    (right). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

    2 Top: center of mass, windward and lee side 50% probability isoline (lines) and power law fits (symbols)

    in the injector midplane. Bottom: Impact of grid resolution: base grid (lines), coarse grid (symbols).

    ρj/ρc = 10 (left) and 100 (right). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

    3 Discrete number (left) and volume (right) frequency distributions of primary atomization drops,

    ρj/ρc = 10 (circles, black) and 100 (squares, gray). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

    4 Location of drop generation side view (center), top view (right), ρj/ρc = 10 (circles, black) and 100

    (squares, gray). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

    5 Primary atomization crossflow direction drop velocities for drop size classes d0 (left) and d4 (right)

    for ρj/ρc = 10 (black) and 100 (gray). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

    6 Primary atomization jet direction drop velocities scaled by jet (top) and crossflow velocity (bottom)

    for drop size classes d0 (left) and d3 (right) for ρj/ρc = 10 (black) and 100 (gray). . . . . . . . . . . . 16

    7 Primary atomization transverse direction drop velocities for drop size classes d0 (left) and d3 (right)

    for ρj/ρc = 10 (black) and 100 (gray). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    19