15
ILASS-Americas 29th Annual Conference on Liquid Atomization and Spray Systems, Atlanta, GA, May 2017 Effects of Sub-Grid Scale (SGS) Dispersion Modeling on Large-Eddy Simulation (LES) of Non-Evaporative and Evaporative Diesel Sprays C.-W. Tsang* and C. J. Rutland Department of Mechanical Engineering University of Wisconsin-Madison 1500 Engineering Drive, Madison, WI 53706, USA Abstract Performance of the sub-grid models accounting for the effects of unresolved motions on Diesel spray dispersion was tested. The models predict the SGS dispersion velocity used for calculating the slip velocity in Lagrangian-Eulerian LES spray models. Discussion was made on the effects of different model formulations and model constants on the simulations of the constant-volume Engine Combustion Network non-evaporative and evaporative “Spray-A”. It was found that the SGS dispersion models have profound impact on the prediction of the spatial distribution of liq- uid mass. Neglecting the SGS dispersion model results in the under-predicted width of the lateral projected liquid mass density profiles. Also, the prediction of the projected liquid mass density is sensitive to the two model con- stants determining the SGS dispersion velocity magnitude and turbulence time scale. On the other hand, the predic- tions of gas-phase velocity profiles, fuel vapor mass fraction profiles, vortex structures, and liquid penetration are insensitive to different SGS dispersion model setups. The primary reason for this is that the motion of high- momentum liquid blobs in the near-nozzle region leading to air entrainment and subsequent gas jet development is little influenced by the SGS dispersion. The SGS dispersion is more critical in determining the motion of droplets having small inertia. *Corresponding author: [email protected]

Effects of Sub-Grid Scale (SGS) Dispersion Modeling on ... · Effects of Sub-Grid Scale (SGS) Dispersion Modeling on Large-Eddy Simulation (LES) of Non-Evaporative and Evaporative

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

  • ILASS-Americas 29th Annual Conference on Liquid Atomization and Spray Systems, Atlanta, GA, May 2017

    Effects of Sub-Grid Scale (SGS) Dispersion Modeling on Large-Eddy Simulation (LES) of Non-Evaporative and Evaporative Diesel Sprays

    C.-W. Tsang* and C. J. Rutland

    Department of Mechanical Engineering University of Wisconsin-Madison

    1500 Engineering Drive, Madison, WI 53706, USA

    Abstract Performance of the sub-grid models accounting for the effects of unresolved motions on Diesel spray dispersion was tested. The models predict the SGS dispersion velocity used for calculating the slip velocity in Lagrangian-Eulerian LES spray models. Discussion was made on the effects of different model formulations and model constants on the simulations of the constant-volume Engine Combustion Network non-evaporative and evaporative “Spray-A”. It was found that the SGS dispersion models have profound impact on the prediction of the spatial distribution of liq-uid mass. Neglecting the SGS dispersion model results in the under-predicted width of the lateral projected liquid mass density profiles. Also, the prediction of the projected liquid mass density is sensitive to the two model con-stants determining the SGS dispersion velocity magnitude and turbulence time scale. On the other hand, the predic-tions of gas-phase velocity profiles, fuel vapor mass fraction profiles, vortex structures, and liquid penetration are insensitive to different SGS dispersion model setups. The primary reason for this is that the motion of high-momentum liquid blobs in the near-nozzle region leading to air entrainment and subsequent gas jet development is little influenced by the SGS dispersion. The SGS dispersion is more critical in determining the motion of droplets having small inertia.

    *Corresponding author: [email protected]

  • Introduction

    In internal combustion engines, fuel injection pro-cess determines the mixing of fuel vapor and air, which greatly impact engine efficiency and emissions. Turbu-lent dispersion of fuel droplets is one of many im-portant physical mechanisms influencing the fuel-air mixing process. The existence of droplets may change the characteristics of turbulent flows, e.g., change of the turbulent kinetic energy spectrum. In turn, a range of length and time scales of turbulent flow may have important effects on the dynamics of droplets. Funda-mental studies on spray-turbulence interaction are ex-tensively carried out by direct numerical simulations (DNS) and experiments in simplified flow configura-tions. Balachandar and Eaton [1] reviewed some im-portant aspects such as preferential concentration of droplets and turbulence modulation from the funda-mental DNS and experimental studies.

    Large-eddy simulation (LES) has become a popu-lar approach in practical engine spray simulations due to the increase of computational power. The major ad-vantage of LES over the conventional Reynolds-averaged Navier-Stokes (RANS) approach is that it can capture more unsteady flow structures, which allows us to study new phenomena such as cycle-to-cycle varia-tions [2]. Unlike DNS resolving flow length scales down to the Kolmogorov scales, LES only resolves filtered flow length scales. This means that the effect of flow length scales below the filter size on spray dy-namics needs to be modelled. This model is commonly called a sub-grid scale (SGS) dispersion model.

    A number of SGS dispersion models have been proposed and tested in non-engine flows. They can be categorized into stochastic and deterministic models. For the stochastic models, Bini and Jones [3] modeled the velocity increment of Lagrangian particles as the Wiener vector process pre-multiplied by a diffusion coefficient matrix, which is a function of the particle-turbulence interaction time and the sub-grid kinetic energy. This model was validated in dilute particle-laden mixing layer [3] and evaporating acetone sprays [4]. Pozorski and Apte [5] modeled the increment of the SGS dispersion velocity itself by the Wiener pro-cess. The coefficients in the stochastic PDE of the SGS dispersion velocity are related to the sub-grid kinetic energy. Preferential concentration of Lagrangian parti-cles was quantified by a radial distribution function (RDF). This model was tested in particle-laden forced isotropic turbulence. It was found that the RDF pre-dicted by this model is in good agreement with DNS data in a large Stokes number case, but it did not work well in a small Stokes number case. Also, the predic-tion is sensitive to the tunable constant determining the time scale of residual motion. One of the deterministic models was proposed by Okong’o and Bellan [6]. The

    SGS dispersion velocity has a magnitude of the filtered standard deviation which is determined from the SGS stress model and the filtered velocity, and its sign is opposite to the Laplacian of the filtered velocity. As-sumptions needed in this model were verified by the DNS data of an evaporating mixing layer laden with evaporating particles in a priori test. Another determin-istic model is to determine the SGS dispersion velocity from the approximate deconvolution of a filtered veloc-ity. The main idea is an approximation of the unfiltered field by truncated series expansion of the inverse filter operator [7,8]. This model was later used by Bharadwaj et al. [9] to calculate the spray source term in the sub-grid kinetic energy transport equation.

    The most commonly used SGS dispersion model for engine spray simulations is inherited from the RANS dispersion model used in the KIVA engine sim-ulation code [10]. The SGS dispersion velocity is sam-pled once every turbulence correlation time from the Gaussian distribution function with variance propor-tional to the sub-grid kinetic energy. This type of mod-el is widely used in LES of engine sprays, e.g. Refs. [11,12]. However, studies on the impacts of the SGS dispersion model on engine spray simulations are few. One of few studies was done by Jangi et al. [13] inves-tigating the effects of the stochastic Gaussian disper-sion model on the LES prediction of Diesel spray. They found that turning off the stochastic dispersion model results in over-predicted liquid penetration and the rate of air entrainment is slower.

    The objective of this study is to understand how different modeling strategies of the SGS dispersion and model parameters influence the prediction of spray-air mixing under Diesel engine operating conditions. The first part of the paper starts with introducing the La-grangian-Eulerian governing equations of liquid and gas flows, followed by the formulation and discussion of the SGS dispersion model. The second part is the validation and evaluation of models performance under non-vaporizing and vaporizing Diesel spray conditions. The paper closes with a summary and conclusions.

    Two-Way Coupled Lagrangian-Eulerian Governing Equations

    A two-way coupled Lagrangian-Eulerian approach is employed to solve the two-phase flow problems. Gas is treated as a continuous phase. The spray is character-ized by a number of discrete computational parcels, with each parcel representing a statistical sample of droplets having the same properties. Properties of each parcel such as position, velocity, and temperature are tracked by solving the Lagrangian governing equations with sub-models predicting physical processes of aero-dynamic drag, atomization, vaporization, etc.

  • The LES gas-phase momentum equation is written as

    𝜕𝜌𝒖𝜕𝑡

    + ∇ ∙ 𝜌𝒖𝒖

    = −∇𝑝 + ∇ ∙ 𝜌𝜈 ∇𝒖 + ∇𝒖 , −23∇ ∙ 𝒖𝑰 − ∇ ∙ 𝜌𝝉

    + 𝑺232

    (1)

    where 𝝉 is the sub-grid stress tensor which is modelled by the mixed dynamic structure model [14]. This mod-el has been shown to have a superior performance over conventional constant coefficient viscosity-based mod-els such as the one-equation model [15]. The momen-tum exchange between the two phases is through the filtered spray momentum source term, 𝑺232, given as

    𝑺232 =1𝑉6

    𝑛8,2𝑚8,2𝒖8,2

    ;<

    2=>

    − 𝑛8,2𝑭8,2

    ;<

    2=>

    (2)

    where the first term on the right-hand side is momen-tum source due to vaporization, and the second term is due to aerodynamic drag force. The symbol 𝑉6 is the filter volume which is taken to be equal to the compu-tational cell volume, 𝑛@ the number of computational parcels in the filter volume, 𝑛8 the number of drops in a parcel, 𝑚8 the vaporization rate of a drop, 𝒖8 the drop velocity, and 𝑭8 the drop drag force.

    The equation of motion for a droplet is written as

    𝑑𝒖8𝑑𝑡

    = 𝜏8C> 𝒖 + 𝒖′ − 𝒖8 (3)

    where the droplet response time scale is

    𝜏8 =431𝐶8𝜌8𝜌𝑑 𝒖 + 𝒖G − 𝒖8 C> (4)

    where 𝐶8 is the drag coefficient, 𝜌8 the droplet density, and 𝑑 the droplet diameter, so the drag force 𝑭8 =𝑚8

    𝑑𝒖𝑑𝑑𝑡

    where 𝑚8 is the droplet mass. Eq. (3) cannot be solved without a model to close the dispersion velocity term, 𝒖′. The model for 𝒖′ is described in the next sec-tion. Note that the SGS dispersion velocity is used not only in the drag calculation, but in atomization, breakup, and vaporization models in which a number of non-dimensional numbers such as Weber number and Reynolds number are based on the slip velocity, 𝒖 + 𝒖′ − 𝒖8. In this section only the momentum conservation equations of gas and liquid phases are shown since the focus of this study is to study the effect of the SGS dispersion on the momentum coupling of liquid and gas. Mass and energy conservation equations of liquid and

    gas are also solved in this study. Their detailed formu-lations can be found in Ref. [14]. SGS Dispersion Models In this study, we assume that the dispersion velocity can be decomposed into the deterministic and the sto-chastic parts,

    𝒖G = 𝒖HIH + 𝒖HJ3. (5)

    The deterministic part, 𝒖HIH, is modeled by the approx-imate deconvolution method [7],

    𝒖HIH = 2𝒖 − 3𝒖 + 𝒖. (6)

    The explicit filtering in Eq. (6) is numerically imple-mented by cell face area-weighted averaging, i.e., 𝒖 =

    𝐴H𝒖HH 𝐴HH and 𝒖 = 𝐴H𝒖HH 𝐴HH where the summation is over all faces (area=𝐴H) of a computa-tional cell. Theoretically, the approximate deconvolu-tion method can only recover the sub-grid scales on the order of the cutoff wave number. Information from smaller scales cannot be recovered, as discussed in Refs [5,6]. In this study, the small-scale motion is rep-resented by the stochastic part of the dispersion veloci-ty. It is assumed to be isotropic and Gaussian-distributed with zero mean:

    𝑓 𝑢HJ3,O =12𝜎Q𝜋

    𝑒𝑥𝑝 −𝑢HJ3,OQ

    2𝜎Q, (7)

    where the subscript i is denoted as the ith component of a vector, and the variance 𝜎Q is proportional to the sub-grid kinetic energy,

    𝜎Q = 𝐶HOI23𝑘HIH (8)

    where 𝐶HOI is a model constant and 𝑘HIH is solved by a transport equation 𝜕𝜌𝑘HIH𝜕𝑡

    + ∇ ∙ 𝜌𝒖𝑘HIH

    = −𝜌𝝉: 𝑺𝒊𝒋 − 𝐶Z𝜌𝑘HIH[ Q

    Δ+ ∇ ∙ 𝜌𝐶]∆𝑘HIH

    >Q ∇𝑘HIH + 𝑊HIH

    (9)

    where 𝑺𝒊𝒋 is the filtered strain rate tensor, 𝐶Z and 𝐶] are model constants, ∆ is the cell size, and the sub-grid kinetic energy spray source term is

    𝑊HIH =1𝑉6

    𝑛8,2𝑭8,2 ∙ 𝒖2G;<

    2=>

    . (10)

  • From Eqs. (2), (3), and (10) we can see that the disper-sion velocity 𝒖′ appears both in the gas- and liquid-phase equations, so the effect of the SGS dispersion is two-way coupled.

    The SGS dispersion velocity of a parcel is sampled once every turbulence correlation time, 𝑡J`ab . This concept was also used in the RANS dispersion model [10]. In RANS, the turbulence correlation time is the minimum of the eddy breakup time and the time for a droplet to traverse a turbulent eddy. The length and the time scales in the RANS model are determined from the turbulent kinetic energy and its dissipation rate. We can extend this model to be used in the LES framework by simply replacing the turbulent kinetic energy by the sub-grid kinetic energy, and the turbulent kinetic ener-gy dissipation rate by the 𝑘HIH dissipation rate, 𝜀HIH , which is modelled as the second term on the right-hand side of Eq. (9) divided by the filtered density. Thus, we can write 𝑡J`ab as

    𝑡J`ab = 𝑚𝑖𝑛𝑘HIH𝜀HIH

    , 𝐶@H𝑘HIH[/Q

    𝜀HIH1

    𝒖 + 𝒖′ − 𝒖8 (11)

    where 𝐶@H = 0.16432 is an empirical constant derived from the RANS 𝑘 − 𝜀 model. Another possible formulation of 𝑡J`ab can be related to the deterministic part of the SGS dispersion velocity,

    𝑡J`ab = 𝐶HIH2∆

    𝒖HIH − 𝒖8 (12)

    where 𝐶HIH is a model constant, 2∆ respresents the smallest eddy LES can resolve. The physical meaning of Eq. (12) is the time for a droplet traveling through the largest unresolved eddy moving with the velocity of 𝒖HIH. The underlying assumption of this formulation is that droplets are smaller than the largest unresolved eddy. Typically, the grid size is on the order of (usually larger than) the nozzle hole diameter in LES Lagrangi-an-Eulerian simulations, so the assumption is reasona-ble. It is of interest to compare individual droplet trajec-tories predicted by Eqs. (11) and (12). Figure 1 shows the trajectories of one of liquid blobs and its child drop-lets due to the Kelvin-Helmholtz breakup under the Engine Combustion Network (ECN) “Spray-A” condi-tions [16]. The trajectories predicted by Eq. (12) are apparently much more reasonable than those predicted by Eq. (11). We can see that in Figure 1(a) all the child droplet trajectories are straight lines. By examining the magnitude of 𝑡J`ab of child droplets predicted by Eq. (11), we found that it is on the order of 10C[ seconds, which is the same order of magnitude of the injection duration. Thus, after a child droplet is born due to the

    Kelvin-Helmholtz breakup with a randomly chosen velocity vector orthogonal to the slip velocity vector of its parent blob, its dispersion velocity almost never gets updated or only gets updated once, resulting in the straight trajectories as shown in Figure 1(a). On the other hand, the turbulence correlation time predicted by Eq. (12) is on the order of 10Ch seconds, so the disper-sion velocity of child droplets does get updated by Eq. (7) multiple times, resulting in the droplet trajectories more similar to the expected stochastic diffusion pro-cess. This comparison suggests that the simple modifica-tion from RANS-type spray models to LES may not work well without consideration of different modeling methodologies between RANS and LES, e.g., turbulent kinetic energy versus sub-grid kinetic energy. In the subsequent study, Eq. (12) is used to calculate the tur-bulence correlation time.

    (a)

    (b)

    Figure 1. Trajectories of one of liquid blobs and its child droplets due to the Kelvin-Helmholtz breakup predicted by (a) Eq. (11) and (b) Eq. (12) with 𝑪𝒔𝒈𝒔 =𝟎. 𝟓. Parent droplets/blobs are shown in solid trajecto-ries, and child droplets are shown in dashed lines.

    Effect of the variance of the Gaussian distribution and the turbulence correlation time on particle trajecto-ries is studied in a simplified case before running full

  • simulations. Consider a particle is located at 𝑧, 𝑦 =0,0 initially, moving with a constant mean speed 𝑈 in

    the z-direction, and the dispersion velocities in z- and y-directions are 𝜁 and 𝜉 respectively. The dispersion velocities are sampled once every constant time inter-val ℎ by the normal distribution with zero mean and a constant variance of 𝜂Q. Assume there is no drag, no collision, and no breakup, then after time T the z- and y-positions of the particle are simply

    𝑧 𝑇 = 𝑈𝑇 + 𝜁Oℎ,/v

    O=>

    (13)

    and

    𝑦 𝑇 = 𝜉Oℎ,/v

    O=>

    . (14)

    The y-position, 𝑦 𝑇 , is also Gaussian distributed with mean zero and variance of

    𝑣𝑎𝑟 𝑦 𝑇 =𝑇ℎℎQ𝜂Q = 𝑇ℎ𝜂Q. (15)

    Hence, larger variance, 𝜂Q, or longer duration between two sampling events, ℎ, lead to larger variance of y-positions and thus larger spreading angle. This is veri-fied by plotting sample paths of particles whose posi-tions are determined by Eqs. (𝟏𝟑) and (14), as shown in Figure 2.

    Figure 2. Particle sample paths calculated from Eqs. (13) and (14) with different variances and time inter-vals between two sampling events.

    Numerical Setups of the ECN Spray-A

    “Spray-A” is a series of spray experiments carried out by Sandia National Laboratories and their collaborators, whose measured data is available from the online database, “Engine Combustion Network (ECN)” [16]. Spray is injected from a single-hole injec-

    tor with a speed of ~500 m/s into a constant-volume pressurized chamber. Its temperature and oxygen con-centration can be controlled. Experimental conditions are listed in Table 1. Note that for the room tempera-ture case, vaporization of spray is negligible, so it is referred to as a non-vaporizing spray. In this study, we are interested in spray behaviors under non-reacting conditions, so the oxygen concentration is zero. The top-hat injection profiles of the non-vaporizing and vaporizing sprays are shown in Figure 3.

    Fuel 100 % n-dodecane Oxygen concentration 0 %

    Ambient gas temperature [K] 900 (vaporizing) and 300 (non-vaporizing)

    Ambient gas density [kg/m3] 22.8 Nozzle hole diameter [mm] (single hole) 0.09

    Nozzle discharge coefficient 0.89 Fuel injection pressure [MPa] 150

    Injection duration [ms] 6 (vaporizing) and 1.5 (non-vaporizing)

    Fuel temperature [K] 373

    Table 1. ECN Spray-A operating conditions.

    Figure 3. Top-hat rate of injection profiles.

    The CFD toolbox, OpenFOAM version 2.1.1

    [17], is used to numerically solve the governing equa-tions of liquid and gas with a number of spray sub-models. Detailed numerical algorithms for solving the partial differential equations can be found in Jasak’s PhD thesis [18]. The authors’ previous work [19] also evaluated the performance of different numerical schemes for solving the gas-phase momentum equation, and found that the central cubic scheme for the convec-tion term with the implicit Euler time integration scheme gave the best prediction.

  • A two-dimensional cut-plane of the three-dimensional domain and grid sizes are shown in Figure 4. Note that all cells are isotropic, namely ∆𝑥 = ∆𝑦 =∆𝑧. The grid size is refined to 0.25 mm in the main spray development region. The computational time step size is fixed at 2.5e-07 seconds, resulting in maximum Courant number of ~0.3. The Euler implicit method is used to numerically integrate the Lagrangian equation of motion, Eq. (3). The time step size for solving this equation is restricted to one-tenth of the droplet re-sponse time scale, 𝜏8 . This constraint improves the numerical stability. It was found that when the variance of the Gaussian distribution, Eq. (8), is large or the turbulence correlation time, Eq. (12), is short the solu-tions are likely to be unstable without this constraint.

    Spray Models

    Liquid blobs are introduced into a computa-tional domain at a given injection position. These blobs experience aerodynamic drag, atomization, and breakup, and smaller droplets are formed. The drag model by Liu et al. [20] takes account for the defor-mation of spherical droplets by correcting the standard drag coefficient. The stochastic Kelvin-Helmholtz/Rayleigh-Taylor (KH-RT) breakup model [19,21] is employed to simulate the atomization and breakup process. This is a modified version of the clas-sical KH-RT model. The model constant determining the KH breakup time scale is determined stochastically and dynamically. For vaporizing sprays, Frossling cor-relation [22] is employed to calculate the vaporization rate of a spherical droplet. Collision and coalescence of droplets are not considered in this study.

    Figure 4. 2-D cut-plane of the computational mesh.

    Test Cases To investigate how different modeling methodologies of SGS dispersion impact the prediction of spray dy-namics, simulations with different model setups as listed in Table 2 were run. Case #1 does not use any dispersion model. Namely, the dispersion velocity 𝒖′ is zero. Case #2 only uses the deterministic model in which 𝒖G is equal to the sub-grid velocity computed from Eq. (6). Cases #3 through #5 are used to study the effect of the magnitude of the variance of the Gaussian distribution, and Cases #3, #6, and #7 are used to study the effect of the turbulence correlation time.

    Case No. Model

    𝑪𝒔𝒊𝒈 in Eq. (8)

    𝑪𝒔𝒈𝒔 in Eq. (12)

    #1 No model 𝒖G = 0

    N/A N/A

    #2 Deterministic 𝒖G = 𝒖𝑠𝑔𝑠

    N/A 0.5

    #3 Deterministic + stochastic 𝒖G = 𝒖HIH + 𝒖HJ3

    0.5 0.5 #4 0.75 0.5 #5 0.25 0.5 #6 0.5 1.0 #7 0.5 0.25

    Table 2. Different model setups to be tested.

    Non-Vaporizing Spray-A Results: Spatial Distribu-tion of Liquid Mass

    The impact of the SGS dispersion model on the spatial distribution of liquid mass is shown in Fig-ure 5, which compares the predicted liquid projected mass density (PMD) profiles against the experimental results from the ECN database. The PMD data were obtained from X-ray radiography path-integrated measurements at Argonne National Laboratory. The data provide a two-dimensional projection of liquid fuel mass onto a plane whose normal vector is orthog-onal to the spray axis. Detailed technique of the meas-urement can be found in Ref. [23]. The data were en-semble-averaged at quasi-steady state at the region of interest. In the simulations, we performed time-averaging (0.4 ms – 1.2 ms) and spatial-averaging (10 different viewing angles) to mimic the ensemble-averaging in the experiments. As shown in Figure 5, without using the stochastic model (Case #1 and #2) the PMDs at y = 0 are significantly over-predicted, and the spray width is under-predicted at downstream re-gion. The stochastic model predicts wider spray and improves the prediction at y = 0. These differences can be explained by Figure 6, showing the mass-weighted and spatially averaged dispersion velocity 𝑢2Z;G and

  • 𝑢a2HG at different axial locations at 0.6 ms. The calcula-tions of 𝑢2Z;G and 𝑢a2HG are

    𝑢2Z;G (𝑍) =𝑚8,2𝑛8,2𝑢2G2=> 𝜒 𝑧8,2, 𝑍

    𝑚8,2𝑛8,22=> (16)

    and

    𝑢a2HG (𝑍) =𝑚8,2𝑛8,2 𝑢2G − 𝑢2Z;G (𝑍)2=>

    Q 𝜒 𝑧8,2, 𝑍𝑚8,2𝑛8,22=>

    (17)

    where

    𝜒 𝑧8, 𝑍 =1, 𝑍 − ∆𝑍 ≤ 𝑧8 ≤ 𝑍 + ∆𝑍0,𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒, (18)

    𝑁 the total number of parcels in the computational do-main, 𝑧8 the z-coordinate of a droplet position, and ∆𝑍 is 0.5 mm. The RMS of the dispersion velocity, shown by the error bars in Figure 6, in the two lateral direc-tions (x and y) predicted by the deterministic model is much smaller than those predicted by the stochastic model. Hence, it is less probable for a droplet moving further away from the centerline, resulting in narrower sprays. Another observation in Figure 6 is that the predicted dispersion velocity is not isotropic upstream of the spray. The mean values in the two lateral direc-tions are approximately zero, and the RMS values in the two lateral directions are close to each other. On the other hand, the mean values in the streamwise direction is positive upstream, and the RMS values is larger than those in the two lateral directions. Since the stochastic part of the model is isotropic, the anisotropy must re-sult from the approximate deconvolution method. The stochastic part of the dispersion model enhances the diffusion of droplets, resulting in better prediction of the PMD profiles. Next, the effect of the model constants, 𝐶HOI and 𝐶HIH, is investigated.

    (a)

    (b)

    Figure 5. Liquid projected mass density profiles at (a) z = 5 mm and (b) z = 10 mm predicted by the different SGS dispersion models.

    (a)

    (b)

    Figure 6. Mass-weighted and spatially averaged dis-persion velocity predicted by (a) Case #2 and (b) Case #3. Symbols are mean values, and error bars are RMS values.

  • Figure 7 shows the qualitative comparison of instantaneous PMD contours predicted by different values of 𝐶HOI and 𝐶HIH. The display range of PMD is from 0 to 1 𝜇𝑔/𝑚𝑚Q. The value of 1 is chosen in that Pickett et al. [24] suggested that spray edge observed from optical diffused backlighting measurements cor-respond to the PMD of 1 𝜇𝑔/𝑚𝑚Q measured from the x-ray measurements. Also, the PMD contours are shown up to z = 20 mm to observe the predictions in the near-nozzle region more closely, although the pene-tration is ~35 mm at 0.55 ms. As shown in Figure 7, larger values of 𝐶HOI or 𝐶HIH predict wider spray. This is consistent with Figure 2 in which the simplified case is considered. Figure 8 and Figure 9 show the quantitative comparison of PMD profiles predicted by different values of 𝐶HOI or 𝐶HIH, respectively. Different values of the two constants predict similar PMD profiles at y = 0.

    Comparing the lateral profiles at two different axial distances, the prediction at z = 10 mm is more sensitive to the two constants than at z = 5 mm. Larger values of 𝐶HOI or 𝐶HIH predict more diffusion of droplets and thus wider profiles. This suggests that the SGS dispersion model has larger impact on the spatial distribution of liquid droplets having small inertia at downstream of the spray.

    Figure 7. Instantaneous PMD contours at 0.55 ms predicted by different values of 𝑪𝒔𝒊𝒈 and 𝑪𝒔𝒈𝒔 in the SGS disper-sion model, Eqs. (8) and (12).

    (a) (b) (c)

    Figure 8. PMD profiles at (a) y = 0 mm, (b) z = 5 mm, (c) z = 10 mm predicted by different values of 𝑪𝒔𝒊𝒈.

  • (a) (b) (c)

    Figure 9. PMD profiles at (a) y = 0 mm, (b) z = 5 mm, (c) z = 10 mm predicted by different values of 𝑪𝒔𝒈𝒔.

    Non-Vaporizing Spray-A Results: Gas-Phase Flow Structures and Statistics

    In the previous section, it is found that the dispersion velocity has profound effect on the spatial distribution of liquid mass. In this section, the effect of the SGS dispersion model on gas-phase solutions are examined. Qualitative comparison is made first by vis-ualizing vortex structures by plotting iso-surfaces of the Q-criterion [25]. The Q-criterion is defined as

    𝑄 =12𝛀Q − 𝑺Q (19)

    where 𝛀 is the rotational rate tensor and S the strain rate tensor. Figure 10 shows the iso-surfaces of the Q-criterion predicted by the different model setups listed

    in Table 2. Unlike the notable difference of the liquid spray angles shown in Figure 7, spreading angles of the gas jet predicted by the different dispersion models are similar to each other. Length scales of the vortices ap-pear to be similar as well. Mean and root-mean-square (RMS) gas veloc-ity radial profiles at three different axial locations are plotted in Figure 11 and Figure 12. Time-averaging from 0.4 ms to 1.2 ms and spatially averaging in the azimuthal direction were performed to obtain mean and RMS values. Although there are some differences of the mean and RMS values at centerline at z = 5 mm and 15 mm, the overall shapes of the profiles predicted by the different model setups are close to each other. These results indicate that the SGS dispersion model plays a less crucial role in predicting the air entrain-ment using the current LES L-E approach under Diesel spray conditions.

    Figure 10. Q-criterion iso-surfaces at 0.4 ms, 𝑸 = 𝟏. 𝟓×𝟏𝟎𝟖𝟏/𝒔𝟐, colored by vorticity magnitude predicted by different model setups. #1: 𝒖′ = 𝟎; #2: 𝒖′ = 𝒖𝒔𝒈𝒔; #5: 𝒖′ = 𝒖𝒔𝒈𝒔 + 𝒖𝒔𝒕𝒐, 𝑪𝒔𝒊𝒈 = 𝟎. 𝟐𝟓; #3:𝒖′ = 𝒖𝒔𝒈𝒔 + 𝒖𝒔𝒕𝒐, 𝑪𝒔𝒊𝒈 =𝟎. 𝟓; #4: 𝒖′ = 𝒖𝒔𝒈𝒔 + 𝒖𝒔𝒕𝒐, 𝑪𝒔𝒊𝒈 = 𝟎. 𝟕𝟓.

  • (a) (b) (c)

    Figure 11. Mean gas velocity profiles predicted by different model setups at three axial locations, (a) 5 mm, (b) 15 mm, and (c) 25 mm. #1: 𝒖′ = 𝟎; #2: 𝒖′ = 𝒖𝒔𝒈𝒔; #5: 𝒖′ = 𝒖𝒔𝒈𝒔 + 𝒖𝒔𝒕𝒐, 𝑪𝒔𝒊𝒈 = 𝟎. 𝟐𝟓; #3:𝒖′ = 𝒖𝒔𝒈𝒔 + 𝒖𝒔𝒕𝒐, 𝑪𝒔𝒊𝒈 =𝟎. 𝟓; #4: 𝒖′ = 𝒖𝒔𝒈𝒔 + 𝒖𝒔𝒕𝒐, 𝑪𝒔𝒊𝒈 = 𝟎. 𝟕𝟓.

    (a) (b) (c)

    Figure 12. Room-mean-square (RMS) gas velocity profiles predicted by different model setups at three different axial locations, (a) 5 mm, (b) 15 mm, and (c) 25 mm. #1: 𝒖′ = 𝟎; #2: 𝒖′ = 𝒖𝒔𝒈𝒔; #5: 𝒖′ = 𝒖𝒔𝒈𝒔 + 𝒖𝒔𝒕𝒐, 𝑪𝒔𝒊𝒈 =𝟎. 𝟐𝟓 ; #3:𝒖′ = 𝒖𝒔𝒈𝒔 + 𝒖𝒔𝒕𝒐, 𝑪𝒔𝒊𝒈 = 𝟎. 𝟓; #4: 𝒖′ = 𝒖𝒔𝒈𝒔 + 𝒖𝒔𝒕𝒐, 𝑪𝒔𝒊𝒈 = 𝟎. 𝟕𝟓.

    This fact is further examined. Figure 13 shows

    time series of the convection, sub-grid stress, pressure gradient, and spray momentum source terms in the gas-phase z-momentum equation, Eq. (2), predicted by Case #5 at different positions. At the upstream region, 5 mm from the injector, the magnitude of the spray momentum source term is comparable to the convec-tion term, meaning a significant amount of momentum transfer from liquid to gas. Figure 14 also shows a sharp decrease of the drag force within 5 mm. Moving further downstream to 10 mm and 15 mm from the injector, the magnitude of the spray momentum source term is much smaller than the convection, sub-grid stress, and pressure gradient terms, and it is negligible at 15 mm. This corresponds to small drag force down-stream as shown in Figure 14. Thus, air entrainment and subsequent gas jet development due to spray injec-tion is mainly initiated in the near-nozzle region (with-in ~5mm). In this region, liquid blobs injected from the

    nozzle carry significant amount of momentum and ex-perience little deceleration as shown in Figure 16. Fig-ure 15 shows that within 5 mm the relative dispersion velocity, 𝒖′ 𝒖 − 𝒖8 , is smaller than that at the downstream region. This suggests that the motion of the liquid blobs leading to air entrainment and subse-quent gas jet development is little influenced by the SGS dispersion. This may be one of the reasons for similar gas velocity profiles predicted by the different dispersion model setups.

    At the downstream region (z > 10 mm), the dispersion velocity magnitude is comparable to the resolved slip velocity as shown in Figure 15, indicating that the dispersion model is important in determining droplets motion. This is further justified by Figure 16, showing that the larger the dispersion velocity magni-tude the larger the acceleration magnitude. This is also consistent with Figure 8(c), indicating that larger dis-persion velocity results in more diffusion of droplets. Another note is that at the downstream region of the

  • spray, it can be approximated as one-way coupled since the momentum coupling term in the gas momentum equation can be negligible as shown in Figure 13(c) and (d), but droplets still experience large accelera-tion/deceleration due to the stochastic SGS dispersion (Figure 16).

    The main point from Figure 13 through Figure 16 is that the SGS dispersion model has small effect on the air entrainment process, but has significant impact on the droplets motion downstream of the spray where momentum transfer from liquid to gas is negligible. In fact, we can clearly see the correlation between the relative dispersion velocity and the drag force in Figure 17: the more important the SGS dispersion the smaller the drag force.

    (a)

    (b)

    (c)

    (d)

    Figure 13 Time series of z-momentum budgets at (a) 𝒙, 𝒚, 𝒛 = 𝟎, 𝟎, 𝟓 , (b) 𝒙, 𝒚, 𝒛 = 𝟎, 𝟎, 𝟏𝟎 , (c) 𝒙, 𝒚, 𝒛 = 𝟎, 𝟎, 𝟏𝟓 , and (d) 𝒙, 𝒚, 𝒛 = 𝟎, 𝟐, 𝟏𝟓 .

    The injection position is 𝒙, 𝒚, 𝒛 = 𝟎, 𝟎, 𝟎 .

    Figure 14. Mass-weighted and spatially averaged drag force magnitude versus distance from injector predicted by the different model setups.

    Figure 15. Mass-weighted and spatially averaged ratio of dispersion velocity magnitude to resolved slip veloc-ity magnitude versus distance from injector predicted by two different values of 𝑪𝒔𝒊𝒈.

  • Figure 16. Mass-weighted and spatially averaged drop-lets acceleration versus distance from injector predicted by the different model setups.

    Figure 17. Drag-relative dispersion velocity scatter plot.

    Vaporizing Spray-A Results

    In this section, simulation results of the vapor-izing Spray-A are compared against the available ex-perimental data. As shown in Figure 18, in average Case #4 (largest 𝐶HOI) predicts the shortest liquid pene-tration, and Case #1 (no model) predicts the longest liquid penetration. This is consistent with the PMD results showing the larger the constant 𝐶HOI the larger the predicted liquid spray angle. Nevertheless, the dif-ference of liquid penetrations predicted by the different model setups is within 1mm, which is small compared to the liquid penetration, ~10mm. This finding is con-sistent with the finding by [26]. Some differences of vapor penetrations predicted by the different model setups exist as shown in Figure 19. Also, vapor pene-tration does not increase or decrease monotonically with the value of 𝐶HOI. The cause of these differences

    might be explained by the different centerline gas ve-locity predictions as shown in Figure 11and Figure 12. In Figure 20, mean fuel vapor mass fraction profiles are compared. To compare the simulation results to the ensemble-averaged experimental data, time-averaging from 1.5 ms to 2.5 ms and spatial averaging in the azi-muthal direction were performed. Except for the cen-terline fuel vapor mass fraction within z = 15 mm, the spatial distributions of fuel vapor mass predicted by the different models and model constants are close to each other, which is consistent with the similar velocity pro-files as shown in Figure 11. From these vaporizing spray results, again we see that the effect of the SGS dispersion modeling on gas entrainment is small.

    Figure 18. Liquid penetration predicted by the differ-ent model setups.

    Figure 19. Vapor penetration predicted by different model setups.

  • (a)

    (b)

    (c)

    Figure 20. Mean fuel vapor mass fraction profiles at (a) centerline, (b) 25 mm, and (c) 40 mm.

    Summary and Conclusions

    This work is focused on the development and evaluation of the SGS dispersion model used in large-eddy simulation of engine sprays. The model computes the dispersion velocity term needed in the calculation of the slip velocity used in a number of spray models. It is assumed that the dispersion velocity is decomposed into the deterministic and the stochastic part. The de-terministic part is calculated by the approximate de-convolution method, which recovers the largest unre-solved scales from the solution of the filtered velocity. Small-scale sub-grid motions are taken account by the stochastic part of the model. The stochastic part of the dispersion velocity is assumed to be isotropic and

    Gaussian distributed. The dispersion velocity is sam-pled once every turbulence correlation time, the time needed for a droplet to traverse an eddy in sub-grid scales. We found that the RANS-type turbulence corre-lation time model predicts unphysical droplet trajecto-ries since the predicted correlation time is on the same order of the injection duration, meaning that the disper-sion velocity almost never gets updated. Thus, a new form of the turbulence correlation time is proposed. It is related to the deterministic part of the model and the filter/computational cell size, Eq. (12).

    The importance of the stochastic part of the model was evaluated. It was found that without the stochastic part, the liquid spray angle is under-predicted. Using the stochastic model enhances the radial diffusion of droplets, resulting in improved predictions of the PMD profiles.

    The effect of the two model constants, 𝐶HOI and 𝐶HIH , was investigated. Larger values of the constant 𝐶HOI and 𝐶HIH predict larger variance of the Gaussian distribution and longer turbulence correlation time, respectively. It was found that larger variance or longer turbulence correlation leads to larger liquid spray angle, which is consistent with the analytical analysis in a simplified case (two-dimensional, no drag, no breakup, and no collision).

    The influence of the SGS dispersion modeling on gas-phase flow structures and statistics was studied. Different dispersion model setups listed in Table 2 pre-dict similar gas jet spreading angle, similar length scales of vortices, and similar mean and RMS values of gas velocity. The reason for this was further investigat-ed. It was found that the SGS dispersion model is more important in determining the motion of droplets having small inertia and thus contributing less to momentum transfer from liquid to gas. On the other hand, for the liquid blobs having large inertia and initiating gas en-trainment and subsequent gas jet development, the dis-persion velocity magnitude is small compared to the resolved slip velocity, so it does not affect the drag force prediction significantly.

    It is suggested that one cannot neglect the SGS dispersion model in LES of engine sprays since it pro-vides more reasonable prediction of the spatial distribu-tion of liquid mass, although the model plays a less important role in predicting gas entrainment due to high-speed spray injection.

    Nomenclature DNS direct numerical simulation ECN Engine Combustion Network KH-RT Kelvin-Helmholtz/Rayleigh-Taylor LES large-eddy simulation PMD projected mass density

  • RANS Reynolds-averaged Navier-Stokes RMS root-mean-square SGS sub-grid scale 𝐶8 drag coefficient 𝐶HOI dispersion model constant, Eq. (8) 𝐶HIH dispersion model constant, Eq. (12) 𝑘HIH sub-grid kinetic energy 𝑚8 droplet mass 𝑛8 number of drops in a parcel 𝑝 filtered pressure 𝑺232 filtered spray momentum source term 𝑡J`ab turbulence correlation time 𝒖 filtered gas velocity 𝒖′ sub-grid dispersion velocity 𝒖8 droplet velocity 𝑉6 filter/computational cell volume 𝑊HIH sub-grid kinetic energy spray source term ∆ computational cell size 𝜀HIH sub-grid kinetic energy dissipation rate 𝜌 filtered density 𝜌8 droplet density 𝜏 sub-grid stress tensor 𝜏8 droplet response time scale 𝜈 molecular kinematic viscosity Acknowledgments This work was supported by Caterpillar Inc. The authors acknowledge additional support from CEI, Inc through their EnSight software used for generating the plots of vortex structures. References [1] Balachandar, S., and Eaton, J. K., 2010,

    “Turbulent dispersed multiphase flow,” Annu. Rev. Fluid Mech., 42(1), pp. 111–133.

    [2] Rutland, C. J., 2011, “Large-eddy simulations for internal combustion engines - a review,” Int. J. Engine Res., 12(5), pp. 421–451.

    [3] Bini, M., and Jones, W. P., 2008, “Large-eddy simulation of particle-laden turbulent flows,” J. Fluid Mech., 614, p. 207.

    [4] Bini, M., and Jones, W. P., 2009, “Large eddy simulation of an evaporating acetone spray,” Int. J. Heat Fluid Flow, 30(3), pp. 471–480.

    [5] Pozorski, J., and Apte, S. V., 2009, “Filtered particle tracking in isotropic turbulence and stochastic modeling of subgrid-scale dispersion,” Int. J. Multiph. Flow, 35(2), pp. 118–128.

    [6] Okong’o, N., and Bellan, J., 2000, “A priori subgrid analysis of temporal mixing layers with

    evaporating droplets,” Phys. Fluids, 12(6), pp. 1573–1591.

    [7] Stolz, S., and Adams, N. A., 1999, “An approximate deconvolution procedure for large-eddy simulation,” Phys. Fluids, 11(1999), pp. 1699–1701.

    [8] Stolz, S., Adams, N. A., and Kleiser, L., 2001, “An approximate deconvolution model for large-eddy simulation with application to incompressible wall-bounded flows,” Phys. Fluids, 13(4), pp. 997–1015.

    [9] Bharadwaj, N., Rutland, C. J., and Chang, S., 2009, “Large eddy simulation modelling of spray-induced turbulence effects,” Int. J. Engine Res., 10(2), pp. 97–119.

    [10] Amsden, A. A., O’Rourke, P. J., and Butler, T. D., 1989, KIVA-II: A computer program for chemically reactive flows with sprays, NM (USA).

    [11] Gong, C., Jangi, M., Lucchini, T., D’Errico, G., and Bai, X.-S., 2014, “Large eddy simulation of air entrainment and mixing in reacting and non-reacting diesel sprays,” Flow, Turbul. Combust., 93(3), pp. 385–404.

    [12] Senecal, P. K., Pomraning, E., and Richards, K. J., 2013, “An investigation of grid convergence for spray simulations using an LES turbulence model,” SAE Tech. Pap., (2013-01–1083).

    [13] Jangi, M., Solsjo, R., Johansson, B., and Bai, X.-S., 2015, “On large eddy simulation of diesel spray for internal combustion engines,” Int. J. Heat Fluid Flow, 53, pp. 68–80.

    [14] Tsang, C.-W., Trujillo, M. F., and Rutland, C. J., 2014, “Large-eddy simulation of shear flows and high-speed vaporizing liquid fuel sprays,” Comput. Fluids, 105, pp. 262–279.

    [15] Yoshizawa, A., and Horiuti, K., 1985, “A statistically-derived subgrid-scale kinetic energy model for the large-eddy simulation of turbulent flows,” J. Phys. Soc. Japan, 54(8), pp. 2834–2839.

    [16] “Engine Combustion Network” [Online]. Available: http://www.sandia.gov/ecn/.

    [17] “OpenFOAM official webpage” [Online]. Available: www.openfoam.com.

    [18] Jasak, H., 1996, “Error analysis and estimation for the finite volume method with applications to fluid flows,” Imperial College London.

    [19] Tsang, C., and Rutland, C., 2016, “Effects of numerical schemes on large eddy simulation of turbulent planar gas jet and diesel spray,” SAE Int. J. Fuels Lubr., 9(1).

    [20] Liu, A. B., Mather, D., and Reitz, R. D., 1993, “Modeling the effects of drop drag and breakup on fuel sprays.,” SAE Technical Paper 930072.

  • [21] Beale, J., and Reitz, R. D., 1999, “Modeling spray atomization with the Kelvin-Helmholtz/Rayleigh-Taylor hybrid model,” At. Sprays, 9(608), pp. 623–650.

    [22] Faeth, G. M., 1977, “Current status of droplet and liquid combustion,” Prog. Energy Combust. Sci., 3(4), pp. 191–224.

    [23] Kastengren, A., Tilocco, F. Z., Duke, D., and Powell, C. F., 2012, “Time-resolved X-Ray radiography of diesel injectors from the Engine Combustion Network,” ICLASS 2012, 12 th Triennial International Conference on Liquid Atomization and Spray Systems, Heidelberg, Germany.

    [24] Pickett, L. M., Manin, J., Kastengren, A., and Powell, C. F., 2014, “Comparison of near-field structure and growth of a diesel spray using light-based optical microscopy and X-ray radiography,” SAE Int. J. Engines, pp. 1044–1053.

    [25] Dubief, Y., and Delcayre, F., 2000, “On coherent-vortex identification in turbulence,” J. Turbul., 1, pp. 1–22.

    [26] Irannejad, A., and Jaberi, F., 2014, “Large eddy simulation of turbulent spray breakup and evaporation,” Int. J. Multiph. Flow, 61, pp. 108–128.