ilass-americas 30th annual conference on liquid atomization … · 2019. 4. 19. · comparison of...

15
ILASS-Americas 30th Annual Conference on Liquid Atomization and Spray Systems, Tempe, AZ, May 2019 _____________________________________ * Corresponding author: [email protected] A Stochastic Kelvin-Helmholtz/Rayleigh-Taylor Breakup Model and a Syn- thetic Eddy Injection Model for Large-Eddy Simulations of Diesel Sprays Chi-Wei Tsang 1,2* , Christopher J. Rutland 1 1 University of Wisconsin-Madison Madison, Wisconsin, 53706, United States 2 The Dow Chemical Company Lake Jackson, TX, 77566, United States Abstract The goal of this study is to develop and improve physical models for large-eddy simulations of Diesel sprays. Particularly, the synthetic eddy injection model and the stochastic Kelvin-Helm- holtz/Rayleigh-Taylor (KH-RT) atomization and breakup model were developed and tested. The synthetic eddy injection model determines fluctuations of velocities of liquid fuel at the spray nozzle exit. The model attempts to simulate turbulence at the nozzle exit without the need of in- ternal nozzle flow simulations by superimposing a number of virtual coherent structures. The idea of the stochastic KH-RT model is to stochastically and dynamically determine several model pa- rameters, specifically the KH and RT length and time scales. This is an attempt to reduce the sensitivity of the model parameters in the standard KH-RT model (Beale, J. and Reitz, R.D., 1999)). The performance of these two newly developed models was compared to the classical models, namely the standard KH-RT and the cone angle injection models (Reitz, R., 1987). Two experimental databases, the Engine Combustion Network constant-volume sprays (Pickett, L. M. et al., 2010) and the Engine Research Center optical engine sprays (Neal, N. & Rothamer, D., 2016) with a range of operating conditions were used to validate the models. The stochastic KH- RT model improved the prediction of the spatial distribution of liquid fuel in the region where secondary breakup occurs. The stochastic KH-RT model was able to better predict liquid penetra- tions in a wider range of operating conditions (errors within 5 %). The synthetic eddy injection model improved the predictions of the spatial distribution of liquid mass in the near-nozzle region and vapor penetrations at early stage of injection. The results suggest that development of insta- bility modes and turbulent transport in the near-nozzle region were better represented by the syn- thetic eddy injection model. The model also showed less computational grid sensitivity. Overall, using these two new models overcomes several limitations in the original models and makes large eddy simulations (LES) as a more predictive tool for Diesel sprays. Beale, J. C., & Reitz, R. D. (1999). Modeling spray atomization with the Kelvin-Helmholtz/Rayleigh-Taylor hybrid model. Atomization and sprays, 9(6). Neal, N., & Rothamer, D. (2016). Measurement and characterization of fully transient diesel fuel jet processes in an optical engine with production injectors. Experiments in Fluids, 57(10), 155. Pickett, L. M., Genzale, C. L., Bruneaux, G., Malbec, L. M., Hermant, L., Christiansen, C., & Schramm, J. (2010). Comparison of diesel spray combustion in different high-temperature, high-pressure facilities. SAE International Journal of Engines, 3(2), 156-181. Reitz, R. (1987). Modeling atomization processes in high-pressure vaporizing sprays. Atomisation and Spray Technology, 3(4), 309-337. .

Upload: others

Post on 31-Jan-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • ILASS-Americas 30th Annual Conference on Liquid Atomization and Spray Systems, Tempe, AZ, May 2019

    _____________________________________

    *Corresponding author: [email protected]

    A Stochastic Kelvin-Helmholtz/Rayleigh-Taylor Breakup Model and a Syn-

    thetic Eddy Injection Model for Large-Eddy Simulations of Diesel Sprays

    Chi-Wei Tsang1,2*, Christopher J. Rutland1

    1University of Wisconsin-Madison

    Madison, Wisconsin, 53706, United States 2The Dow Chemical Company

    Lake Jackson, TX, 77566, United States

    Abstract

    The goal of this study is to develop and improve physical models for large-eddy simulations of

    Diesel sprays. Particularly, the synthetic eddy injection model and the stochastic Kelvin-Helm-

    holtz/Rayleigh-Taylor (KH-RT) atomization and breakup model were developed and tested. The

    synthetic eddy injection model determines fluctuations of velocities of liquid fuel at the spray

    nozzle exit. The model attempts to simulate turbulence at the nozzle exit without the need of in-

    ternal nozzle flow simulations by superimposing a number of virtual coherent structures. The idea

    of the stochastic KH-RT model is to stochastically and dynamically determine several model pa-

    rameters, specifically the KH and RT length and time scales. This is an attempt to reduce the

    sensitivity of the model parameters in the standard KH-RT model (Beale, J. and Reitz, R.D.,

    1999)). The performance of these two newly developed models was compared to the classical

    models, namely the standard KH-RT and the cone angle injection models (Reitz, R., 1987). Two

    experimental databases, the Engine Combustion Network constant-volume sprays (Pickett, L. M.

    et al., 2010) and the Engine Research Center optical engine sprays (Neal, N. & Rothamer, D.,

    2016) with a range of operating conditions were used to validate the models. The stochastic KH-

    RT model improved the prediction of the spatial distribution of liquid fuel in the region where

    secondary breakup occurs. The stochastic KH-RT model was able to better predict liquid penetra-

    tions in a wider range of operating conditions (errors within 5 %). The synthetic eddy injection

    model improved the predictions of the spatial distribution of liquid mass in the near-nozzle region

    and vapor penetrations at early stage of injection. The results suggest that development of insta-

    bility modes and turbulent transport in the near-nozzle region were better represented by the syn-

    thetic eddy injection model. The model also showed less computational grid sensitivity. Overall,

    using these two new models overcomes several limitations in the original models and makes large

    eddy simulations (LES) as a more predictive tool for Diesel sprays.

    Beale, J. C., & Reitz, R. D. (1999). Modeling spray atomization with the Kelvin-Helmholtz/Rayleigh-Taylor

    hybrid model. Atomization and sprays, 9(6). Neal, N., & Rothamer, D. (2016). Measurement and characterization of fully transient diesel fuel jet processes in

    an optical engine with production injectors. Experiments in Fluids, 57(10), 155.

    Pickett, L. M., Genzale, C. L., Bruneaux, G., Malbec, L. M., Hermant, L., Christiansen, C., & Schramm, J. (2010).

    Comparison of diesel spray combustion in different high-temperature, high-pressure facilities. SAE International

    Journal of Engines, 3(2), 156-181.

    Reitz, R. (1987). Modeling atomization processes in high-pressure vaporizing sprays. Atomisation and Spray

    Technology, 3(4), 309-337.

    .

  • ILASS-Americas 30th Annual Conference on Liquid Atomization and Spray Systems, Tempe, AZ, May 2019

    2

    Introduction

    In recent years, large-eddy simulation (LES) has be-

    come a promising approach to simulate unsteady fuel

    spray dynamics [1]. In LES, filtered turbulent flow fields

    are directly solved on computational grids. The charac-

    teristic of the filtered flow fields is that large-scale struc-

    tures of turbulent flow are retained, while small-scale

    structures are filtered out. Mathematically, the filtered

    turbulent flow field is obtained by a convolution opera-

    tion [2]. As pointed out by Rutland [1], there are several

    expectations of LES applied in engine simulations. The

    primary expectation is that more vortices, eddies, and

    flow structures can be resolved. That is, LES provides

    the “large eddy” flow field. More predicted flow struc-

    tures might imply that LES is more suitable to be used to

    study more complex phenomena due to turbulence, such

    as cycle-to-cycle variability and engine knock. To com-

    putationally study these phenomena, accurate prediction

    of spatially and temporally evolving turbulent flows is

    required. Generally, LES is a more appropriate approach

    than RANS for these types of studies. However, LES for

    engine sprays is still a relatively new research field com-

    pared to RANS and a major issue is discussed in the fol-

    lowing.

    There are few sub-grid spray models specifically for

    LES. Spray droplets are usually several orders of magni-

    tude smaller than the computational cell sizes. In LES,

    even in direct numerical simulations (DNS) of engine

    sprays, it is unlikely to resolve gas phase flow field near

    droplet surfaces [3]. Therefore, sub-grid models ac-

    counting for interactions of liquid droplets, sub-grid gas

    motion, and resolved gas motion may be critical. When

    moving towards LES of sprays, the most common ap-

    proach is to use LES sub-grid models developed for sin-

    gle-phase gas flow, accompanied with spray sub-models

    developed and tuned in the RANS framework [1]. Often

    only simple extensions are made to the RANS models,

    including replacing turbulent kinetic energy and ensem-

    ble-averaged velocity by sub-grid kinetic energy and fil-

    tered velocity, respectively. This approach can provide

    reasonable predictions, e.g., [4], [5], in terms of global

    quantities such as spray tip penetrations in some particu-

    lar operating conditions. However, as shown in later sec-

    tions in this paper, these simple extensions may fail in

    many other operating conditions and are not able to give

    reasonable predictions of all quantities of interest. A pos-

    sible cause is that the RANS Lagrangian spray models

    were developed to interact with ensemble-averaged gas

    phase models, without considerations of unsteady fluc-

    tuations and interactions of different flow length scales.

    In addition, the RANS-based spray models have a num-

    ber of model constants. These model constants have been

    optimized with RANS simulations. However, these opti-

    mized constants may not be suitable for LES. Spray

    models taking the unique properties of LES into account

    are few. To meet the LES expectations mentioned above,

    we should continue to explore possible next-generation

    LES spray approaches.

    The objective of this study is to address this issue by

    developing spray submodels consistent with the LES

    framework. Particularly, a new atomization and breakup

    model, the stochastic KH-RT model, and a new injection

    model, the synthetic eddy injection model, for LES of

    Diesel sprays were developed. The synthetic eddy injec-

    tion model dynamically determines initial velocities of

    Lagrangian parcels by superposition of virtual coherent

    structures. The stochastic KH-RT model is based on the

    classical KH-RT hybrid model but with stochastic time

    and length scales of the two breakup mechanisms.

    Lagrangian-Eulerian approach

    The present study uses the discrete parcel Lagran-

    gian-Eulerian approach to simulate two-phase flow

    problems [6]. Droplets having the same properties are

    grouped into a parcel. Based on the assumption that the

    parcel does not occupy any volume with respect to the

    gas phase, the parcel can be treated as point sources of

    mass, momentum, and energy for the gas phase.

    The present study employs the dynamic structure

    sub-grid stress tensor model [7] with the addition of a

    near nozzle viscosity model to simulate gas phase turbu-

    lence. The purpose of adding the near nozzle viscosity

    model is to enhance sub-grid mixing in the near-nozzle

    region where a portion of energy-containing motions

    may not be adequately resolved under moderate grid res-

    olution [8].

    Injection

    In this study, the blob injection method is used [9].

    This method assumes that large “blobs” exit the nozzle

    position with a diameter equal to the nozzle hole diame-

    ter. In addition to the blob size and position, velocity of

    a blob must be given as initial conditions to solve the

    Lagrangian governing equations. In this section, two dif-

    ferent methods to determine the initial velocity are intro-

    duced.

    Conventional Cone Angle Injection -- In RANS

    spray modeling, the spray injection cone angle is speci-

    fied. If the injector axis is parallel to the z-axis, the three

    components of the injection velocity, 𝑉𝑖𝑛𝑗, are given by

    𝑉𝑖𝑛𝑗,𝑧 = 𝑉0,

    𝑉𝑖𝑛𝑗,𝑥 = 𝑉0𝑡𝑎𝑛𝜃

    2𝑐𝑜𝑠𝜙,

    𝑉𝑖𝑛𝑗,𝑦 = 𝑉0𝑡𝑎𝑛𝜃

    2𝑠𝑖𝑛𝜙,

    (1)

    where 𝑉0 is given by a measured rate of injection profile, 𝜃 is assumed to be uniformly distributed between 0 and 𝜃𝑚𝑎𝑥, and 𝜙 is assumed to be uniformly distributed be-tween 0 and 2𝜋. In this study, 𝜃𝑚𝑎𝑥 is the cone angle and is chosen as 15 (deg). This value is within the range of

  • 3

    Diesel spray angle measurements in different operating

    conditions [10].

    The parcel distribution predicted by this specified

    spray cone angle method with LES is shown in Error!

    Reference source not found.. It can be seen that the

    cone angle method predicts the dispersion of spray in an

    ensemble-averaged sense, although the gas-phase was

    predicted by LES. This is not consistent with the ap-

    proach of LES. The goal of LES is to capture unsteady

    and three-dimensional structures of the spray. In experi-

    mental [11], [12] and high-fidelity CFD images [13] of

    sprays, large-scale oscillations of sprays are clearly ob-

    served in the near-nozzle region, suggesting that flow

    with different wavenumber modes are developed due to

    shear instability and/or turbulence inside the nozzle. The

    spray angle method perturbs the injection velocity ran-

    domly without any time or space correlation. This white-

    noise perturbation helps little in terms of developing a

    range of wavenumber modes which nonlinearly interact

    with each other to enhance turbulent mixing.

    Figure 1. Liquid parcels distribution predicted by

    the cone angle injection method with LES.

    Synthetic Eddy Injection -- The objective of de-veloping a new injection model is to simulate turbulence

    at the nozzle outlet and to help LES develop instability

    modes not only through the basic equations solved on the

    computational mesh but also by artificially perturbing

    the injection velocity.

    The method adopted here is called the synthetic

    eddy method originally developed by Jarrin et al. [14].

    This method was used extensively in generating bound-

    ary conditions for LES or DNS of single-phase flows.

    Here, we employ this method to generate initial velocity

    data for Lagrangian parcels. The injection velocity, 𝑉𝑖𝑛𝑗,

    is decomposed into mean and fluctuating parts,

    𝑉𝑖𝑛𝑗,𝑖 = �̅�𝑖 + 𝑎𝑖𝑗𝑣𝑗′ , (2)

    where the tensor 𝑎𝑖𝑗 is obtained from a prescribed Reyn-

    olds stress tensor. Here, the Reynolds stresses are for the

    flow inside the injector. If the injector axis is parallel to

    the z-axis, and the injection position is fixed at the center

    of the nozzle outlet, and axisymmetric flow is assumed,

    then only the isotropic components of the Reynolds

    stress tensor are non-zero at the centerline of the injector.

    Thus, the injection velocity can be written as

    𝑉𝑖𝑛𝑗,𝑧 = 𝑉0 + √𝑅𝑧𝑧𝑣𝑧′ ,

    𝑉𝑖𝑛𝑗,𝑥 = √𝑅𝑥𝑥𝑣𝑥′ ,

    (3)

    𝑉𝑖𝑛𝑗,𝑦 = √𝑅𝑦𝑦𝑣𝑦′ ,

    where 𝑉0 is given by a measured rate of injection profile. The fluctuating signal, 𝑣′, is determined by superposition of coherent structures. Error! Reference source not

    found. helps explain the idea of the model for 𝑣′. There are N eddies at positions

    𝑝1(𝑡), 𝑝2(𝑡), … , 𝑝𝑘(𝑡), … , 𝑝𝑁−1(𝑡), 𝑝𝑁(𝑡) in a cylinder whose center is located at the injection position, 𝑝𝑖𝑛𝑗 =

    (𝑥𝑖𝑛𝑗 , 𝑦𝑖𝑛𝑗 , 𝑧𝑖𝑛𝑗), length is 2𝑙𝑠, and diameter is 2(𝑙𝑠 +

    𝑟𝑛𝑜𝑧) where 𝑙𝑠 is a prescribed length scale. The velocity at 𝑝𝑖𝑛𝑗 is the sum of the contributions from the N eddies.

    Mathematically, this is written as

    𝑣𝑗′(𝑡)

    =1

    √𝑁∑ 𝜀𝑘𝑗

    𝑁

    𝑘=1

    √𝑉𝐶𝑙𝑠

    3𝑓 (

    𝑥𝑖𝑛𝑗 − 𝑥𝑘(𝑡)

    𝑙𝑠) 𝑓 (

    𝑦𝑖𝑛𝑗 − 𝑦𝑘(𝑡)

    𝑙𝑠)

    ∗ 𝑓 (𝑧𝑖𝑛𝑗 − 𝑧𝑘(𝑡)

    𝑙𝑠)

    (4)

    where the index 𝑗 can be for x, y, or z, 𝑉𝐶 is the cylinder volume, and the shape function 𝑓 is given by

    𝑓(𝑏) = {√3

    2(1 − |𝑏|) 𝑖𝑓 |𝑏| < 1

    0, 𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒.

    (5)

    The function 𝑓 satisfies the normalized condition

    ∫ 𝑓2(𝑏)𝑑𝑏1

    −1

    = 1 (6)

    so that 𝑣𝑗′ has zero mean and unit variance. Positions of

    the 𝑁 turbulent eddies are chosen randomly at 𝑡 = 0 within the cylinder, and these eddies are transported with

    the constant speed of 𝑉0 in the z-direction. At time 𝑡 +∆𝑡, the position of an eddy is thus updated by

    𝑧𝑘(𝑡 + ∆𝑡) = 𝑧𝑘(𝑡) + 𝑉0∆𝑡 𝑥𝑘(𝑡 + ∆𝑡) = 𝑥𝑘(𝑡) 𝑦𝑘(𝑡 + ∆𝑡) = 𝑦𝑘(𝑡),

    (7)

    if 𝑧𝑘(𝑡) < 𝑧𝑖𝑛𝑗 + 𝑙𝑠, i.e. the eddy at time 𝑡 is still inside

    the cylinder. If 𝑧𝑘(𝑡) ≥ 𝑧𝑖𝑛𝑗 + 𝑙𝑠, the eddy is recycled

    back to the inlet face of the cylinder, i.e., 𝑧𝑘(𝑡 + ∆𝑡) =𝑧𝑖𝑛𝑗 − 𝑙𝑠, and 𝑥𝑘(𝑡 + ∆𝑡) and 𝑦𝑘(𝑡 + 𝑑𝑡) are chosen ran-

    domly within the inlet face. The symbol 𝜀𝑘𝑗 in Eq. (4) is

    the sign of eddy k on component j. Its value is either 1 or

    -1, and it is chosen randomly with equal probability

    when 𝑡 = 0 and when the eddy 𝑘 is regenerated on the inlet face of the cylinder. The number of eddies, N,

    should be chosen to ensure that inlet plane is statistically

    covered with synthetic eddies. In the current study, N is

    taken to be 100.

    The two prescribed model parameters, the Reynolds

    stresses tensor, 𝑅𝑖𝑗, and the length scale, 𝑙𝑠, were deter-

    mined using data from high-fidelity VoF simulations

    [13], [15]. Values of these two parameters are

    √𝑅𝑧𝑧 = 0.024𝑉0,

    √𝑅𝑥𝑥 = √𝑅𝑦𝑦 = 0.019𝑉0, (8)

  • 4

    and

    𝑙𝑠 = 11𝑟𝑛𝑜𝑧 . (9) Note that the value of the length scale, 11𝑟𝑛𝑜𝑧 , is

    larger than the grid size in practical Diesel spray simula-

    tions. For example, in ECN Spray-A simulations, the

    typical grid size is 0.25 mm, and the length scale 𝑙𝑠 =11𝑟𝑛𝑜𝑧 = 11 ∗ 0.045 = 0.495 𝑚𝑚. This means that the synthetic structures can be discretized by the LES grid.

    Jarrin et al. [16] also suggested that one of the criteria for

    choosing the length scale is that the length scale should

    be at least larger than the grid size.

    Using the synthetic eddy injection method in an LES

    simulation, the parcels distribution as shown in Error!

    Reference source not found. are significantly different

    from that predicted by the spray angle method as shown

    in Error! Reference source not found.. Large-scale liq-

    uid structures can be seen indicating that this new injec-

    tion model is more consistent with LES than the original

    RANS based cone-angle model. More quantitative com-

    parisons including gas-phase flow structures are in later

    sections.

    Figure 2. Schematic concept of the synthetic eddy

    model.

    Figure 3. Liquid parcels distribution predicted by

    the synthetic eddy injection method.

    Atomization and Breakup

    The hybrid Kelvin-Helmholtz/Rayleigh-Taylor

    (KH-RT) breakup model developed two decades ago

    [17] has been widely used in industry and implemented

    into a number of commercial CFD codes. Formulation

    and limitations of this model and an improved version

    are discussed in the following sections.

    Standard KH-RT -- The KH breakup and the RT

    breakup stand for two different breakup mechanisms.

    The KH breakup predicts the wave growth at a gas-liquid

    interphase due to the interaction of aerodynamic drag,

    surface tension force, and liquid inertia. The RT breakup

    mechanism accounts for wave on a droplet induced by

    acceleration normal to the interface between two fluids

    of different densities. The most unstable wave length and

    frequency of each of these two mechanisms can be de-

    rived analytically by performing linear stability analysis

    [9], [18]. Then the length and the time scales are incor-

    porated into the model parameters used to determine the

    droplet size after breakup and the breakup time. Detailed

    procedure of the model can be found in (Beale and Reitz

    1999).

    The KH-RT model is widely used in RANS simula-

    tions, and in recent years it has been employed in LES

    spray simulations, e.g. [19]–[21]. A common issue of the

    KH-RT model used in both RANS and LES is that sim-

    ulation results heavily depend on the choice of the model

    constants. As pointed out by Reitz [9] and Beale and

    Reitz [17], the choice of the value of the KH time con-

    stant, 𝐵1, depends on nozzle flow conditions, but there is no guidance of how to relate the value of the constant to

    the nozzle flow conditions. The model constants may

    also depend on fuel. Brakora et al. [22] showed that to

    improve the prediction of biodiesel sprays, the values of

    𝐵1, the RT size constant, 𝐶𝑅𝑇, and the breakup length constant, 𝐶𝑏, need to be adjusted from the standard by a factor of 2. The choice of the value of 𝐵1 may also vary with grid resolution. Senecal et al. [4] used a value of 7

    for 𝐵1 in a very refined mesh (∆𝑥 = 0.0625 𝑚𝑚). Fu-jimoto et al. [23] studied the effect of 𝐵1 on the predic-tion of spray tip penetrations and droplet size, and found

    that larger value of 𝐵1 predicts longer liquid penetration and larger Sauter mean diameter (SMD). Tsang and Rut-

    land [24] studied the effect of KH size constant, 𝐵0, Kh time constant, 𝐵1, and breakup length constant, 𝐶𝑏. It was found that the momentum exchange between liquid

    and gas is sensitive to 𝐵1. Smaller value of 𝐵1 predicts larger momentum transfer from liquid to gas, resulting in

    longer vapor penetration. It was also found that the pre-

    diction of liquid penetration is sensitive to the constant,

    𝐶𝑏. Smaller 𝐶𝑏 predicts shorter liquid penetration. Perini and Reitz [25] performed the multi-objective optimiza-

    tion to find the best set of the model constants which can

    give reasonable predictions of liquid penetration, vapor

    penetration, and vapor mass distribution under the En-

    gine Combustion Network “Spray-A” conditions [26]. It

  • 5

    was found that a value giving good prediction of an ob-

    jective is not an optimal value for another objective. For

    instance, an optimal value of the RT time constant, 𝐶𝜏, to predict vapor penetration and fuel vapor distribution well

    is ~0.1, but the optimal value for liquid penetration is

    2.39.

    These studies suggest that the KH-RT model con-

    stants are not universal. An optimal set of the model con-

    stants may only be effective for a specific flow configu-

    ration, a specific mesh resolution, or a specific simula-

    tion output. This statement is further justified in the sim-

    ulation results shown in this paper. Thus, the model pre-

    dictive capability is limited. The KH-RT model only pre-

    dicts atomization and breakup due to wave instability.

    Other possible breakup mechanisms such as turbulence

    [27] and nozzle cavitation [28], [29] are neglected. This

    may be one of the reasons why predictions are highly

    sensitive to the model constants.

    Stochastic KH-RT -- The central idea of the sto-

    chastic KH-RT model is that the model parameters are

    determined stochastically and dynamically according to

    local flow conditions. The model parameters include i)

    KH breakup time, 𝜏𝐾𝐻, ii) Breakup length, 𝐿𝑏, iii) RT breakup time, 𝜏𝑅𝑇 , and iv) RT wavelength, 𝜆𝑅𝑇 . That is, even if two parcels have the exact same conditions (We-

    ber number, Ohnesorge number, normal acceleration,

    etc.), they may still have different breakup characteris-

    tics due to the stochastic model. This is an attempt to re-

    duce the sensitivity of the model constants partly result-

    ing from the incapability of predicting breakup mecha-

    nisms other than wave instabilities

    The four parameters mentioned above are all func-

    tions of the droplet radius. Instead of using a fixed value

    of the droplet radius, the droplet radius in the equations

    of the four parameters are determined stochastically by

    time-dependent probability density functions (PDFs).

    The KH breakup time is written as

    𝜏𝐾𝐻,𝑠𝑡𝑜 = 3.726𝐵1𝑟𝑑,𝑠𝑡𝑜𝐾𝐻 𝛬𝐾𝐻𝛺𝐾𝐻⁄ , (10) where 𝛺𝐾𝐻 is the fastest KH growth rate, and 𝛬𝐾𝐻 is the corresponding wavelength, derived from linear stability

    analysis. The rate equation for the parent droplet size, 𝑟𝑑, becomes

    𝑑𝑟𝑑𝑑𝑡

    = −(𝑟𝑑 − 𝑎) 𝜏𝐾𝐻,𝑠𝑡𝑜⁄ , (11)

    where 𝑎 is the child droplet radius assumed to be linearly proportional to 𝛬𝐾𝐻. By assuming inviscid flow and in-finite Weber number, the breakup length can be written

    as [17]

    𝐿𝑏,𝑠𝑡𝑜 = 𝑈𝑟𝑒𝑙𝜏𝐾𝐻,𝑠𝑡𝑜

    =1

    √𝜋𝐵1

    𝑟𝑑,𝑠𝑡𝑜𝐾𝐻𝑟𝑑

    √𝐴𝑛𝑜𝑧𝜌𝑑𝜌𝑐

    , (12)

    where 𝑈𝑟𝑒𝑙 is the relative velocity magnitude between liquid and gas, 𝐴𝑛𝑜𝑧 the nozzle cross-sectional area, 𝜌𝑑 the droplet density, and 𝜌𝑐 the gas density. The form of

    the PDF for 𝑟𝑑,𝑠𝑡𝑜𝐾𝐻 refers to the stochastic breakup model developed by Apte et al. [30]. It follows Kolmo-

    gorov’s hypothesis, which is that the probability to break

    each parent particle into a given number of fragments is

    independent of the parent particle size, and a differential

    Fokker-Planck equation for the PDF of droplet radii can

    be derived. The solution of the differential equation is a

    log-normal distribution function in the form of, 𝑓𝐾𝐻(𝑟𝑑,𝑠𝑡𝑜𝐾𝐻 , 𝑡𝑟,𝐾𝐻)=

    1

    𝑟𝑑,𝑠𝑡𝑜𝐾𝐻√2𝜋〈𝜉𝐾𝐻2 〉𝜈𝐾𝐻𝑡𝑟,𝐾𝐻

    𝑒𝑥𝑝 [−(𝑙𝑛𝑟𝑑,𝑠𝑡𝑜𝐾𝐻 − 𝑙𝑛𝑟𝑑,0 − 〈𝜉𝐾𝐻〉𝜈𝐾𝐻𝑡𝑟,𝐾𝐻)

    2

    2〈𝜉𝐾𝐻2 〉𝜈𝐾𝐻𝑡𝑟,𝐾𝐻

    ],

    (13)

    if 𝑡𝑟,𝐾𝐻 > 0. If 𝑡𝑟,𝐾𝐻 = 0, 𝑟𝑑,𝑠𝑡𝑜𝐾𝐻 = 𝑟𝑑. The KH breakup frequency, 𝜈𝐾𝐻 , is the reciprocal of the stochas-tic KH time, 𝜏𝐾𝐻,𝑠𝑡𝑜. The characteristic time, 𝑡𝑟,𝐾𝐻, for a parcel is tracked and is set or reset to zero if certain con-

    ditions are met. More details of the conditions can be

    found in [31].

    The two shape parameters,〈𝜉𝐾𝐻2 〉𝜈𝐾𝐻𝑡𝑟,𝐾𝐻 and

    〈𝜉𝐾𝐻〉𝜈𝐾𝐻𝑡𝑟,𝐾𝐻 in Eq. (13), determine the shape of the distribution. By definition, 〈𝜉𝐾𝐻〉 and 〈𝜉𝐾𝐻

    2 〉 are the first two moments of 𝜉𝐾𝐻 where 𝜉𝐾𝐻 is the logarithm of the ratio of the size of a child droplet to the size of a parent

    droplet. These two parameters are determined dynami-

    cally according to local flow conditions and fuel proper-

    ties. Apte et al. came up with an assumption and related

    the droplets breakup process to Einstein’s theory of

    Brownian motion to determine these two parameters dy-

    namically. The assumption is that, in the intermediate

    range of scales between large droplets with large Weber

    numbers and the maximum stable droplets with a critical

    Weber number, there is no preferred length scale. This is

    analogous to the inertial range of homogenous turbu-

    lence. Thus, by assuming the relationship of

    𝑈𝑟𝑒𝑙3

    Λ𝐾𝐻 ~

    𝑈𝑟𝑒𝑙,𝑐𝑟3

    Λ𝐾𝐻,𝑐𝑟, (14)

    one can obtain

    𝑟𝑑,𝑐𝑟𝑟𝑑

    ~ (Λ𝐾𝐻

    Λ𝐾𝐻,𝑐𝑟)

    2/3

    (𝑊𝑒𝑐𝑟𝑊𝑒𝑔

    ), (15)

    where 𝑈𝑟𝑒𝑙,𝑐𝑟 is the magnitude of relative velocity at which disruptive forces are balanced by capillary forces

    (similar to the turbulent velocity scale of the smallest ed-

    dies), 𝑟𝑑,𝑐𝑟 = 𝑊𝑒𝑐𝑟𝜎 𝜌𝑐𝑈𝑟𝑒𝑙2⁄ is the maximum stable

    droplet radius where 𝑊𝑒𝑐𝑟 is the critical Weber number taken to be a value of 6, 𝜎 the surface tension coefficient, Λ𝐾𝐻,𝑐𝑟 is the critical KH wavelength written as

    𝛬𝐾𝐻,𝑐𝑟

    =9.02𝑟𝑑(1 + 0.45𝑂ℎ

    0.5)(1 + 0.4𝑇𝑎𝑐𝑟0.7)

    (1 + 0.87𝑊𝑒𝑐𝑟1.67)0.6

    (16)

  • 6

    where 𝑇𝑎𝑐𝑟 = 𝑂ℎ𝑊𝑒𝑐𝑟0.5 where 𝑂ℎ is the Ohnesorge

    number. Taking the logarithm of Eq. (15), the model for

    〈𝜉𝐾𝐻〉 can be written as 〈𝜉𝐾𝐻〉

    =2

    3𝑙𝑛 (

    Λ𝐾𝐻Λ𝐾𝐻,𝑐𝑟

    ) + 𝐾1,𝐾𝐻𝑙𝑛 (𝑊𝑒𝑐𝑟𝑊𝑒𝑔

    ), (17)

    where 𝐾1,𝐾𝐻 is a model constant. The model for the sec-ond moment of 𝜉𝐾𝐻 proposed by Apte et al. based on Einstein’s theory of Brownian motion is written as

    〈𝜉𝐾𝐻2 〉 = −

    〈𝜉𝐾𝐻〉

    𝐾2,𝐾𝐻𝑙𝑛 (𝑟𝑑

    𝑟𝑑,𝑐𝑟)

    (18)

    where 𝐾2,𝐾𝐻 is another model constant.

    The stochastic RT breakup time is written as

    𝜏𝑅𝑇,𝑠𝑡𝑜 =𝐶𝜏

    𝛺𝑅𝑇,𝑠𝑡𝑜 (19)

    where 𝛺𝑅𝑇,𝑠𝑡𝑜 is given by

    𝛺𝑅𝑇,𝑠𝑡𝑜 = √2

    3√3𝜎

    [−𝑔𝑡,𝑠𝑡𝑜(𝜌𝑑 − 𝜌𝑐)]3 2⁄

    𝜌𝑑 + 𝜌𝑐 (20)

    where

    𝑔𝑡,𝑠𝑡𝑜 = 𝑔𝑡 (𝑟𝑑

    𝑟𝑑,𝑠𝑡𝑜𝑅𝑇)

    3

    . (21)

    The stochastic RT wavelength is given by

    𝜆𝑅𝑇,𝑠𝑡𝑜 =2𝜋𝐶𝑅𝑇𝐾𝑅𝑇,𝑠𝑡𝑜

    , (22)

    where the stochastic RT wave number is

    𝐾𝑅𝑇,𝑠𝑡𝑜 = √−𝑔𝑡,𝑠𝑡𝑜(𝜌𝑑 − 𝜌𝑐)

    3𝜎. (23)

    Analogous to the log-normal distribution for

    𝑟𝑑,𝑠𝑡𝑜𝐾𝐻, the stochastic radius, 𝑟𝑑,𝑠𝑡𝑜𝑅𝑇 , is also randomly chosen from the log-normal distribution,

    𝑓𝑅𝑇(𝑟𝑑,𝑠𝑡𝑜𝑅𝑇 , 𝑡𝑟,𝑅𝑇)= 1

    𝑟𝑑,𝑠𝑡𝑜𝑅𝑇√2𝜋〈𝜉𝑅𝑇2 〉𝜈𝑅𝑇𝑡𝑟,𝑅𝑇

    ∗ 𝑒𝑥𝑝 [−(𝑙𝑛𝑟𝑑,𝑠𝑡𝑜𝑅𝑇 − 𝑙𝑛𝑟𝑑,0 − 〈𝜉𝑅𝑇〉𝜈𝑅𝑇𝑡𝑟,𝑅𝑇)

    2

    2〈𝜉𝑅𝑇2 〉𝜈𝑅𝑇𝑡𝑟,𝑅𝑇

    ],

    (24)

    if 𝑡𝑟,𝑅𝑇 > 0. If 𝑡𝑟,𝑅𝑇 = 0, 𝑟𝑑,𝑠𝑡𝑜𝑅𝑇 = 𝑟𝑑. The RT breakup frequency, 𝜈𝑅𝑇 , is the reciprocal of the stochastic RT breakup time, 𝜏𝑅𝑇,𝑠𝑡𝑜. Analogous to the KH shape pa-rameters, the RT shaper parameters, 〈𝜉𝑅𝑇〉 and 〈𝜉𝑅𝑇

    2 〉 are respectively given by

    〈𝜉𝑅𝑇〉 =2

    3𝑙𝑛 (

    𝜆𝑅𝑇𝜆𝑅𝑇,𝑐𝑟

    ) + 𝐾1,𝑅𝑇𝑙𝑛 (𝑊𝑒𝑐𝑟𝑊𝑒𝑔

    ), (25)

    and

    〈𝜉𝑅𝑇2 〉 = −

    〈𝜉𝑅𝑇〉

    𝐾2,𝑅𝑇𝑙𝑛 (𝑟𝑑

    𝑟𝑑,𝑐𝑟)

    , (26)

    where the critical RT wavelength is given by [18]

    𝜆𝑅𝑇,𝑐𝑟 = 2𝜋 (𝜎

    −𝑔𝑡,𝑠𝑡𝑜(𝜌𝑑 − 𝜌𝑐))

    1 2⁄

    , (27)

    and 𝐾1,𝑅𝑇 and 𝐾2,𝑅𝑇 are model constants. The model constants used in the standard and stochas-

    tic KH-RT models for the remainder of the current study

    are listed in Table 1 and Table 2. Note that although there

    are 4 more model constants of the stochastic model, this

    set of constants is expected to be more universal and are

    not changed from one operating condition to another.

    Table 1. Standard KH-RT model constants.

    KH size constant, 𝐵0 0.61

    KH time constant, 𝐵1 40

    RT size constant, 𝐶𝑅𝑇 0.1

    RT time constant, 𝐶𝜏 1.0 Breakup length constant, 𝐶𝑏 2.0

    Table 2. Stochastic KH-RT model constants.

    KH size constant, 𝐵0 0.61

    KH time constant, 𝐵1 80

    RT size constant, 𝐶𝑅𝑇 0.1

    RT time constant, 𝐶𝜏 1.0

    𝐾1,𝐾𝐻 1.0

    𝐾2,𝐾𝐻 0.007

    𝐾1,𝑅𝑇 1.0

    𝐾2,𝑅𝑇 0.008

    Results and Discussion

    The injection and atomization and breakup models

    were systematically validated against two experimental

    datasets, the Engine Combustion Network (ECN) sprays

    [26] and the ERC optical engine sprays [32]. A range of

    operating conditions are used for model validation, as

    shown in Table 3. The ERC optical engine conditions are

    listed in Table 4. Four cases using different combinations

    of the breakup and the injection models are evaluated, as

    listed in Table 5.

  • 7

    Table 3. Selected experimental conditions of ECN sprays for model validations.

    Case name Temperature

    (K)

    Density

    (kg/m3)

    Nozzle diameter

    (mm)

    Injection

    pressure

    (MPa)

    Fuel Oxygen (%)

    Non-vap A 300 22.8 0.09 150 C12H26 0

    Vap A 900 22.8 0.09 150 C12H26 0

    Vap A,

    High T and

    low den.

    1400 7.6 0.09 150 C12H26 0

    Vap A, low

    inj. P 900 22.8 0.09 50 C12H26 0

    Vap A,

    low T 700 22.8 0.09 150 C12H26 0

    Vap H 1000 14.8 0.1 150 C7H16 0

    Table 4. ERC optical engine specification.

    Engine type Single-cylinder optically accessible

    Injector Bosch Crin 2 six-hole

    Fuel type #2 Diesel

    Bore (mm) 82.55

    Stroke (mm) 76.2

    Bowl depth (mm) 6.12

    Bowl diameter (mm) 65.18

    Clearance (mm) 1.04

    Compression ratio 14.03

    Engine speed (rpm) 1200

    Cylinder wall temperature (℃) 50 Start of injection (SOI) timing -5 ATDC

    Injection duration (𝜇s) 1500 Injector nozzle diameter (mm) 0.11

    Angle of the plume off the horizontal plane (degs) 13

    Mean temperature at ROI (°C) 945.1±11.4

    Table 5. Test matrix to evaluate breakup and injection

    models.

    Atomization and

    breakup

    Injection

    Case 1 Standard KH-RT Cone angle

    Case 2 Stochastic KH-RT Cone angle

    Case 3 Standard KH-RT Synthetic eddy

    Case 4 Stochastic KH-RT Synthetic eddy

    Simulated spatial distributions of liquid mass are com-

    pared against the X-ray ECN data [33] for non-vaporiz-

    ing sprays (Non-vap A in Table 3) as shown in Figure 4

    to Figure 6. Mesh size was 0.25 mm. The experimental

    results were time- and ensemble-averaged. Simulation

    results were time-averaged and spatial-averaged in the

    azimuthal direction. The duration of the time-averaging,

    0.4 ms to 1.2 ms ASOI, is the same in the simulations

    and in the experiments. Comparing Case 1 and Case 2,

    the standard KH-RT and the stochastic KH-RT models

    have similar predictions in the upstream region. Both

    cases under predict the PMDs at y = 0 before z = 6 mm.

    Moving further downstream, a notable difference be-

    tween the PMD profiles at y = 0 predicted by the stand-

    ard KH-RT and those predicted by the stochastic KH-RT

    can be seen. The experimental data shows that the PMD

    at y = 0 decreases with increased axial distance, as shown

    in Figure 4, but the standard KH-RT model was not able

    to capture this trend. The PMDs at y = 0 predicted by the

    standard KH-RT model remain almost constant beyond

    z ≈ 6 mm. This issue was also reported in the work done by [34] using the same injection model and the breakup

  • 8

    model as Case 1. On the other hand, the stochastic KH-

    RT model was able to capture the trend although the

    PMDs were still over-predicted downstream.

    Comparing Case 1 and Case 3, Case 1 predicted larger

    spray angle than that predicted by Case 3, as shown in

    the contour plots. Also, the injection model plays a more

    important role in the prediction of the PMDs upstream.

    Before 6 mm, the synthetic eddy model gives much bet-

    ter prediction than the cone angle injection model. In

    fact, the PMD prediction is sensitive to the prescribed

    cone angle in the cone angle injection model. Smaller

    prescribed cone angle predicts higher PMD at y = 0, and

    vice versa. It was found that reducing the prescribed cone

    angle from 15° (current value) to 10° did improve the PMD prediction in the upstream region. However, the

    PMD prediction beyond 𝑧 ≈ 6 𝑚𝑚 was even more over-predicted, and the liquid penetration in the vaporizing

    sprays was over-predicted as well. Thus, simply adjust-

    ing the prescribed cone angle cannot improve the predic-

    tions of multiple quantities simultaneously.

    As shown in Figure 6(a), the synthetic eddy model im-

    proves the lateral PMD profile upstream (z = 5 mm). In

    the downstream region at z = 10 mm, the width of the

    lateral profile is under-predicted (Figure 6(b)) by Case 4,

    suggesting that the spray is less dispersed. In fact, the

    prediction of the width of the lateral PMD profiles in the

    downstream region not only depends on the breakup and

    the injection model but the SGS dispersion model. More

    discussion on the effect of the SGS dispersion model on

    the prediction of PMD profiles can be found in [35].

    Nevertheless, overall Case 4 significantly improves the

    prediction of the spatial distribution of liquid mass. The

    synthetic eddy injection model improves the prediction

    in the upstream region, and the stochastic KH-RT model

    improves the prediction in the downstream region.

    Figure 4. Mean liquid projected mass density

    (PMD) contour plots predicted by Case 1, Case 2, Case

    3, and Case 4, and measured by the ECN X-ray non-va-

    porizing spray experiment.

    Figure 5. Comparison of the mean PMD profiles at

    y = 0 predicted by the four cases against the ECN X-ray

    data.

  • 9

    (a)

    (b)

    Figure 6. Mean lateral PMD profiles at (a) z =

    5mm and (b) z = 10 mm. The line labels are the same as

    those in Figure 5.

    The next dataset is the ECN vaporing spray A. Liquid

    and vapor penetrations are compared. The performance

    of the models in predicting the quasi-steady liquid pene-

    trations under different operating conditions of the ECN

    vaporizing sprays can be quantified by computing the er-

    ror defined as

    𝑒𝑟𝑟𝑜𝑟 (%) =�̅�𝑠𝑖𝑚 − �̅�𝑒𝑥𝑝

    �̅�𝑒𝑥𝑝× 100% (28)

    where the overbar means the time-averaging performed

    once the liquid penetration becomes quasi-steady. The

    errors are shown in Figure 7. The experimental quasi-

    steady liquid penetration values are also shown in the

    figure. Compared to the standard vaporizing Spray A

    case, the high temperature and low density case shows

    longer liquid penetration. This is because the drag force

    is smaller due to the lower air density although the air

    carries higher energy to vaporize the fuel. The lower

    temperature case has the longest penetration among

    these five operating conditions due to the slower vapori-

    zation rate. The penetration of the lower injection pres-

    sure case is close to that of the standard case. This occurs

    because both air entrainment rate and liquid fuel flow

    rate depend linearly on the injection velocity [36]. As a

    result, the lower injection pressure and higher injection

    pressure cases require the same liquid length needed to

    entrain enough air to vaporize the fuel [37]. The Spray-

    H penetration is slightly shorter than that of the standard

    Spray-A case since the Spray-H fuel, n-heptane, has

    lower heat of vaporization than the Spray-A fuel, n-do-

    decane, and the ambient temperature is higher in Spray-

    H. Droplet size distribution would affect air entrainment

    rate and vaporization rate which are important factors in

    determining the liquid length, so a good atomization and

    breakup model is needed to predict reasonable liquid

    lengths.

    In general, the errors from the stochastic KH-RT

    model are much smaller than those from the standard

    KH-RT model. Except for the high temperature and low

    density case, the errors from Case 2 and Case 4 are less

    than 2%. On the other hand, Case 1 using the standard

    KH-RT over predicted the liquid penetrations by more

    than 25 % in the low temperature and the low injection

    pressure cases. To achieve better predictions of these two

    cases using the standard KH-RT model, it is unavoidable

    to tune the breakup model constants to enhance breakup

    rate and thus reduce the liquid penetrations. Error! Ref-

    erence source not found. shows that the stochastic KH-

    RT model is more capable of predicting the liquid pene-

    trations reasonably well in a range of operating condi-

    tions without tuning the model constants.

    Figure 7. Errors of the quasi-steady liquid penetra-

    tion predictions of the ECN vaporizing sprays under

    different operating conditions: (a) Spray A, (b) Spray A

    high T and low Den, (c) Spray A low T, (d) Spray A

    low inj. P, and (e) Spray H. The horizontal labels repre-

    sent the different model setups listed in Table 5. The

    experimental values of the quasi-steady liquid penetra-

    tions are also shown in the figures.

  • 10

    Error! Reference source not found. shows the “Vap

    A” mass-weighted and spatially averaged mean and root-

    mean-square (RMS) values of the droplet diameter ver-

    sus distance from the injector. Three distinct slopes can

    be seen in Figure 8. The three distinct slopes imply dif-

    ferent breakup mechanisms from upstream to down-

    stream of the spray. In the near-nozzle region, the droplet

    size decreases gradually with increased distance. The

    KH breakup is the primary mechanism in this region. In

    the breakup model, parent parcels become smaller, and

    child droplets are formed with the diameter proportional

    to the KH wavelength. The RMS values of the droplet

    diameter are larger predicted by the stochastic model

    since the KH breakup time used in the rate equation (Eq.

    (11)) is stochastically sampled from the log-normal dis-

    tribution. The mean values of the droplet diameter start

    to decrease more dramatically at certain axial distances

    which are identified as the breakup length. Beyond the

    breakup length, the RT breakup, which tends to produce

    small drops, is activated. As shown in the figure, the

    breakup length of the “Vap A” case is approximately

    5𝑚𝑚. In this secondary breakup region, the decrease in the mean droplet size with increasing distance predicted

    by the stochastic model is more gradual than that pre-

    dicted by the standard model, and the RMS values are

    larger. The more gradual decrease is due to the stochastic

    breakup length (Eq. (12)), which is determined not only

    by the fluid density ratio but the PDF with the shape pa-

    rameters dependent on local flow conditions. The larger

    RMS values of the droplet size are due to the stochastic

    RT wavelength and the stochastic RT breakup time. Near

    the end of the liquid sprays, the droplet size becomes

    more stable since in this region the liquid slip velocity is

    not large enough to further break up drops through either

    of the mechanisms. In general, the stochastic KH-RT

    model, attempting to account for breakup mechanisms

    other than the primary wave instabilities such as turbu-

    lence, predicts a wider range of the droplet size distribu-

    tion due to the stochastic model parameters. This may be

    one reason why the stochastic model gives better predic-

    tions of the liquid penetrations in a range of operating

    conditions (see Figure 7).

    Figure 8. “Vap A” mass-weighted and spatially

    averaged mean and RMS values of the droplet diameter

    versus distance at 1.5 ms.

    Figure 9 compares the predicted vapor penetrations

    against the ECN experimental data under the different

    operating conditions detailed in Table 3. The vapor pen-

    etration in simulations is defined as the maximum dis-

    tance from an injector where the fuel vapor mass fraction

    is 0.1 %. As shown in the figures, the injection model has

    bigger effect than the breakup model on predicting the

    vapor penetrations, particularly at early times. Within

    0.25 ms, the tip penetration speeds (slopes of the vapor

    penetrations) are over predicted by the cone angle injec-

    tion model. The issue of the over-predicted vapor pene-

    trations at early times (within 1 ms) in LES of engine

    sprays has been reported in many studies [4], [20], [38]–

    [42]. The primary cause is that the turbulent diffusion

    due to interactions of different instability modes is not

    well-captured. One solution is to increase grid resolution

    with higher-order numerical schemes so that flow struc-

    tures in the near-nozzle region can be better-resolved,

    leading to an earlier development of the turbulent diffu-

    sion and shorter vapor penetrations. However, reducing

    grid size by 1/2 in each side of cells may increase com-

    putational time by approximately 16 times (number of

    cells increases by 8 times and the time step size needs to

    reduce by 1/2 to keep the same Courant number). To

    keep a reasonable computational cost while having rea-

    sonable predictions of the vapor penetration, the syn-

    thetic eddy injection model is an alternative. As shown

    in the figures, the synthetic eddy injection model im-

    proves the predictions of the tip penetration speeds at

    early times of fuel injection. The model helps develop

    flow structures faster by artificially providing time- and

    space-correlated injection velocity data.

  • 11

    (a)

    (b)

    (c)

    (d)

    Figure 9. Comparison of the (a) “Vap A”, (b) “Vap

    A high T and low den”, (c) “Vap A low inj P”, and (d)

    “Vap H” vapor penetrations predicted by the different

    spray models against the ECN experimental data. The

    shaded area is measured mean uncertainty.

    Figure 10 compares the vapor penetrations predicted

    by Case 1 and Case 4 using the different mesh resolu-

    tions. The vertical dashed lines in these figures associate

    with the times at which the tip penetration speeds start to

    decrease due to the formation of small-scale flow struc-

    tures downstream of the jet. The vertical lines move to

    the left as the mesh is refined, which is expected since

    flow structures are developed faster with the finer

    meshes. Also, with the 0.25 mm and the 0.5 mm meshes,

    Case 4 predicted earlier developments of flow structures

    and improved the predictions of the vapor penetrations.

    As shown in Figure 11 of fuel vapor contours, more fine

    structures can be seen in the 0.125 mm mesh, as ex-

    pected. For the coarser meshes, 0.5 mm and 0.25 mm,

    Case 4 predicts larger jet angle, suggesting that turbulent

    diffusion is better represented. Overall, the vapor pene-

    trations predicted by Case 4 are less grid-sensitive.

  • 12

    (a)

    (b)

    Figure 10. Vapor penetrations of the “Vap A” case

    predicted by (a) Case 1 and (b) Case 4 using different

    grid resolutions. The penetration speeds reduce at cer-

    tain times indicated by the blue dashed lines, due to the

    development of flow structures downstream of the jet.

    Figure 11. Fuel vapor contour plots of the “Vap

    A” case at 1.0 ms predicted by Case 1 and Case 4 using

    the three different grid resolutions, 0.5 mm, 0.25 mm,

    and 0.125 mm.

    Figure 12 compares side views of the fuel vapor contours

    of one of the six plumes in the ERC optical engine. The

    fuel vapor mass distributions predicted by the standard

    KH-RT model (Case 1 and Case 3) and the stochastic

    KH-RT model (Case 2 and Case 4) are remarkably dif-

    ferent. At 139 𝜇𝑠, the standard KH-RT model predicts two vapor regions. There are gaps with very low fuel va-

    por mass fractions between the two regions with higher

    mass fractions. The location of the vapor region down-

    stream corresponds to the deterministic breakup length

    (~ 10 mm). Beyond the breakup length, vaporization rate

    is faster due to the RT breakup. In the stochastic KH-RT

    model transition from the primary breakup to the second-

    ary breakup does not occur at a fixed breakup length, re-

    sulting in the smoother centerline mass fraction profiles.

    Also, the stochastic KH-RT model predicts higher mass

    fractions in the near-nozzle region.

  • 13

    Figure 12. Fuel vapor mass fraction contours of the injection pressure 52 MPa and ambient density 15 kg/m3 case

    predicted by Cases 1, 2, 3, and 4 at three different times after start of injection. The locations of the small vapor regions

    predicted by the standard KH-RT model at 𝟏𝟑𝟗 𝝁𝒔 correspond to the location where the RT breakup is activated.

    Figure 13 compares the predicted liquid and vapor

    penetrations against the experimental data. The meas-

    ured liquid penetration becomes quasi-steady at approx-

    imately 250 𝜇𝑠. This time was accurately predicted by Case 2 and Case 4 using the stochastic KH-RT model.

    This good prediction also means that the air entrainment

    due to the fuel injection was captured reasonably well.

    On the other hand, the standard KH-RT model predicted

    overshoots of the liquid penetrations. In addition, the liq-

    uid penetrations predicted by Case1 show larger standard

    deviations, indicating that the predictions are more grid-

    sensitive. Overall, Case 4 gives the best prediction of the

    liquid penetration. The vapor penetrations predicted by

    Case 2 and Case 4 before 150 𝜇𝑠 match well with the data. The sharp increases in the vapor penetrations pre-

    dicted by the standard KH-RT model from ~5 mm at 100

    𝜇𝑠 to ~10 mm at 139 𝜇𝑠 correspond to the small vapor regions that formed downstream as shown in the contour

    plots in Figure 12.

    (a)

    (b)

    Figure 13. (a) Liquid and (b) vapor penetrations of

    the injection pressure 52 MPa and ambient density 15

    kg/m3case. The symbols and the error bars represent

    means and standard deviations of the liquid and vapor

    penetrations from the six plumes. The shaded area is the

    experimental 95 % confidence interval band.

    Summary and Conclusions

    Two new LES specific Diesel spray models were

    developed. They are the synthetic eddy injection model

    and the stochastic KH-RT breakup model. Key ideas and

    features of the two models are summarized as follows.

    The synthetic eddy injection model provides time-

    and space-correlated initial velocities for injected La-

    grangian parcels. The model simulates turbulence at the

    nozzle outlet and helps LES develop instability modes

    not only through the basic equations solved on the com-

    putational mesh but also by artificially, but coherently,

    perturbing the injection velocity. The injection velocity

    is modeled by the superposition of coherent structures

    characterized by shape functions dependent on the struc-

    ture positions and a prescribed length scale. The two

    model parameters, the Reynolds stresses and the length

    scale, were derived from the high-fidelity VoF spray

    simulations.

    The stochastic KH-RT model predicts the atomiza-

    tion and breakup processes of Diesel sprays. The model

    attempts to reduce the sensitivity of the model constants

    in the original KH-RT model. The sensitivity partly re-

    sults from the limitation of predicting breakup mecha-

    nisms only from wave instabilities. Compared to the

    original KH-RT model, the stochastic KH-RT model de-

    termines time and length scales of the growth of instabil-

    ity waves stochastically and dynamically. All these

    scales are functions of the stochastic droplet size, which

    is sampled from a log-normal distributions with shape

    parameters determined from local flow conditions.

    The models were systematically tested using the two

    experimental datasets, the ECN constant-volume sprays

    database and the ERC optical engine sprays database. It

    was found that the predictions of near-nozzle flow fields

    are sensitive to the spray injection model. In the non-va-

    porizing Spray-A case, the prediction of the near-nozzle

    liquid PMD profiles was greatly improved by the syn-

    thetic eddy injection model. The cone-angle injection

    model significantly under-predicted the PMDs at y = 0

    in the upstream region. In the vaporizing ECN spray

    cases, the synthetic eddy injection model improved the

    vapor penetration predictions at early times of spray in-

    jection under a range of operating conditions. Also, it

  • 14

    was found that the gas-phase solutions predicted by the

    synthetic eddy injection model are less sensitive to grid

    resolutions.

    The major advantage of the stochastic KH-RT

    model over the standard KH-RT model is that it is able

    to give reasonable predictions of the liquid penetrations

    in a wide range of operating conditions without tuning

    the model constants. The stochastic KH-RT model can

    correctly capture the liquid penetration trends varying

    with injection pressures, ambient densities, and ambient

    temperatures. Results showed that the stochastic KH-RT

    model predicted a wider range of the droplet sizes and a

    smoother transition from the primary KH breakup to the

    secondary RT breakup.

    Overall, compared to the commonly used cone-an-

    gle injection and the standard KH-RT breakup models,

    the synthetic eddy model and the stochastic KH-RT

    breakup model together improve the LES predictability

    of Diesel sprays. Quantities such as liquid penetration,

    vapor penetration, spatial distribution of liquid fuel and

    fuel vapor were well-captured.

    References

    [1] C. J. Rutland, “Large-eddy simulations for

    internal combustion engines - a review,” Int. J.

    Engine Res., vol. 12, no. 5, pp. 421–451, 2011.

    [2] S. B. Pope, Turbulent Flows. Cambridge

    University Press, 2000.

    [3] X. Jiang, G. A. Siamas, K. Jagus, and T. G.

    Karayiannis, “Physical modelling and advanced

    simulations of gas–liquid two-phase jet flows in

    atomization and sprays,” Prog. Energy Combust.

    Sci., vol. 36, no. 2, pp. 131–167, 2010.

    [4] P. K. Senecal, E. Pomraning, and K. J. Richards,

    “An Investigation of Grid Convergence for

    Spray Simulations using an LES Turbulence

    Model,” SAE Tech. Pap., no. 2013-01-1083,

    2013.

    [5] C. Gong, M. Jangi, T. Lucchini, G. D’Errico, and

    X.-S. Bai, “Large eddy simulation of air

    entrainment and mixing in reacting and non-

    reacting diesel sprays,” Flow, Turbul. Combust.,

    vol. 93, no. 3, pp. 385–404, 2014.

    [6] A. A. Amsden, P. J. O’Rourke, and T. D. Butler,

    “KIVA-II: A computer program for chemically

    reactive flows with sprays,” 1989.

    [7] E. Pomraning and C. J. Rutland, “Dynamic One-

    equation Nonviscosity Large-eddy Simulation

    Model,” AIAA J., vol. 40, no. 4, pp. 689–701,

    2002.

    [8] C.-W. Tsang, M. F. Trujillo, and C. J. Rutland,

    “Large-eddy simulation of shear flows and high-

    speed vaporizing liquid fuel sprays,” Comput.

    Fluids, vol. 105, pp. 262–279, 2014.

    [9] R. D. Reitz, “Modeling Atomization Processes

    in High-Pressure Vaporizing Sprays,” At.

    Sprays, vol. 3, pp. 309–337, 1987.

    [10] J. D. Naber and D. L. Siebers, “Effects of Gas

    Density and Vaporization on Penetration and

    Dispersion of Diesel Sprays,” in SAE Paper

    960034, 1996.

    [11] H. G. Fujimoto, H. Adachi, T. Yano, N.

    Marubayashi, T. Hori, and J. Senda, “Spatial

    Droplets Size Distribution in a Diesel Spray

    Taken by Photography with Super High

    Resolution,” ILASS - Eur. 2010, 23rd Annu.

    Conf. Liq. At. Spray Syst., no. September, pp. 1–

    6, 2010.

    [12] Z.-M. Cao, K. Nishino, S. Mizuno, and K. Torii,

    “PIV measurement of internal structure of diesel

    fuel spray,” Exp. Fluids, vol. 29, no. 7, pp. S211-

    -S219, 2000.

    [13] S. S. Deshpande, S. R. Gurjar, and M. F. Trujillo,

    “A computational study of an atomizing liquid

    sheet,” Phys. Fluids, vol. 27, no. 8, 2015.

    [14] N. Jarrin, S. Benhamadouche, D. Laurence, and

    R. Prosser, “A synthetic-eddy-method for

    generating inflow conditions for large-eddy

    simulations,” Int. J. Heat Fluid Flow, vol. 27,

    no. 4, pp. 585–593, 2006.

    [15] S. R. Gurjar, “A Computational Study of

    Injected Liquid Sheets and Jets,” University of

    Wisconsin-Madison M.S. Thesis, 2016.

    [16] N. Jarrin, R. Prosser, J. C. Uribe, S.

    Benhamadouche, and D. Laurence,

    “Reconstruction of turbulent fluctuations for

    hybrid RANS/LES simulations using a

    Synthetic-Eddy Method,” Int. J. Heat Fluid

    Flow, vol. 30, no. 3, pp. 435–442, 2009.

    [17] J. Beale and R. D. Reitz, “Modeling spray

    atomization with the Kelvin-

    Helmholtz/Rayleigh-Taylor hybrid model,” At.

    Sprays, vol. 9, no. 608, pp. 623–650, 1999.

    [18] R. Bellman and R. H. Pennington, “Effects of

    Surface Tension and Viscosity on Taylor

    Instability,” Q. Appl. Math., vol. 12, no. 2, pp.

    151–162, 1953.

    [19] S. Banerjee and C. J. Rutland, “Study on Spray

    Induced Turbulence Using Large Eddy

    Simulations,” At. Sprays, vol. 25, no. 4, pp. 285–

    316, 2015.

    [20] L. Zhou, K. H. Luo, W. Qin, M. Jia, and S. J.

    Shuai, “Large eddy simulation of spray and

    combustion characteristics with realistic

    chemistry and high-order numerical scheme

    under diesel engine-like conditions,” Energy

    Convers. Manag., vol. 93, pp. 377–387, 2015.

    [21] Y. Pei et al., “Large eddy simulation of a

    reacting spray flame with multiple realizations

    under compression ignition engine conditions,”

  • 15

    Combust. Flame, vol. 162, no. 12, pp. 4442–

    4455, 2015.

    [22] J. L. Brakora, Y. Ra, and R. D. Reitz,

    “Combustion Model for Biodiesel-Fueled

    Engine Simulations using Realistic Chemistry

    and Physical Properties,” SAE Int. J. Engines,

    vol. 4, no. 1, pp. 931–947, 2011.

    [23] H. Fujimoto, T. Hori, and J. Senda, “Effect of

    Breakup Model on Diesel Spray Structure

    Simulated by Large Eddy Simulation,” SAE

    Tech. Pap., no. 2009-24-0024, 2009.

    [24] C. Tsang and C. J. Rutland, “Effects of Fuel

    Physical Properties and Breakup Model

    Constants on Large-eddy Simulation of Diesel

    Sprays,” in Proceedings of the ASME 2015

    Internal Combustion Engine Division Fall

    Technical Conference, 2015.

    [25] F. Perini and R. D. Reitz, “Improved

    atomization, collision and sub-grid scale

    momentum coupling models for transient

    vaporizing engine sprays,” Int. J. Multiph. Flow,

    vol. 79, pp. 107–123, 2016.

    [26] “Engine Combustion Network.” [Online].

    Available: http://www.sandia.gov/ecn/.

    [27] Y. Liao and D. Lucas, “A literature review of

    theoretical models for drop and bubble breakup

    in turbulent dispersions,” Chem. Eng. Sci., vol.

    64, no. 15, pp. 3389–3406, 2009.

    [28] J.-S. Han, P.-H. Lu, X.-B. Xie, M.-C. Lai, and N.

    a. Henein, “Investigation of Diesel Spray

    Primary Break-up and Development for

    Different Nozzle Geometries,” in SAE Technical

    Paper 2002-01-2775, 2002.

    [29] H. K. Suh and C. S. Lee, “Effect of cavitation in

    nozzle orifice on the diesel fuel atomization

    characteristics,” Int. J. Heat Fluid Flow, vol. 29,

    no. 4, pp. 1001–1009, 2008.

    [30] S. V Apte, M. Gorokhovski, and P. Moin, “LES

    of atomizing spray with stochastic modeling of

    secondary breakup,” Int. J. Multiph. Flow, vol.

    29, no. 9, pp. 1503–1522, 2003.

    [31] C. Tsang, “Large-Eddy Simulation Modeling of

    Diesel Sprays,” University of Wisconsin

    Madison PhD Thesis, 2017.

    [32] N. Neal and D. Rothamer, “Measurement and

    characterization of fully transient diesel fuel jet

    processes in an optical engine with production

    injectors,” Exp. Fluids, vol. 57, no. 10, p. 155,

    2016.

    [33] A. Kastengren, F. Z. Tilocco, D. Duke, and C. F.

    Powell, “Time-resolved X-Ray radiography of

    diesel injectors from the Engine Combustion

    Network,” in ICLASS 2012, 12 th Triennial

    International Conference on Liquid Atomization

    and Spray Systems, Heidelberg, Germany, 2012.

    [34] Q. Xue et al., “An Eulerian CFD model and X-

    ray radiography for coupled nozzle flow and

    spray in internal combustion engines,” Int. J.

    Multiph. Flow, vol. 70, pp. 77–88, 2015.

    [35] C. W. Tsang, C. W. Kuo, M. Trujillo, and C.

    Rutland, “Evaluation and validation of large-

    eddy simulation sub-grid spray dispersion

    models using high-fidelity volume-of-fluid

    simulation data and engine combustion network

    experimental data,” Int. J. Engine Res., 2018.

    [36] D. L. Siebers, “Liquid-Phase Fuel Penetration in

    Diesel Sprays,” in SAE Technical Paper 980809,

    1998.

    [37] D. L. Siebers, “Scaling liquid-phase fuel

    penetration in diesel sprays based on mixing-

    limited vaporization,” in SAE Technical Paper

    1999-01-0528, 1999.

    [38] T. Hori et al., “Large Eddy Simulation of Non-

    Evaporative and Evaporative Diesel Spray in

    Constant Volume Vessel by Use of KIVALES,”

    in SAE Technical Paper 2006-01-3334, 2006.

    [39] N. Bharadwaj, C. J. Rutland, and S. Chang,

    “Large eddy simulation modelling of spray-

    induced turbulence effects,” Int. J. Engine Res.,

    vol. 10, no. 2, pp. 97–119, 2009.

    [40] A. Wehrfritz, V. Vuorinen, O. Kaario, and M.

    Larmi, “A High Resolution Study of Non-

    Reacting Fuel Sprays using Large-Eddy

    Simulations,” in ICLASS 2012, 12th Triennial

    International Conference on Liquid Atomization

    and Spray Systems, 2012.

    [41] A. Irannejad and F. Jaberi, “Large eddy

    simulation of turbulent spray breakup and

    evaporation,” Int. J. Multiph. Flow, vol. 61, pp.

    108–128, 2014.

    [42] M. Jangi, R. Solsjo, B. Johansson, and X.-S. Bai,

    “On large eddy simulation of diesel spray for

    internal combustion engines,” Int. J. Heat Fluid

    Flow, vol. 53, pp. 68–80, 2015.