A numerical simulation of soot formation in spray flames

Similar documents
Spray evaporation model sensitivities

Large-eddy simulation of an industrial furnace with a cross-flow-jet combustion system

Soot formation in turbulent non premixed flames

Budget analysis and model-assessment of the flamelet-formulation: Application to a reacting jet-in-cross-flow

Lecture 8 Laminar Diffusion Flames: Diffusion Flamelet Theory

Large-eddy simulation of an evaporating and reacting spray

DNS of droplet evaporation and combustion in a swirling combustor

Unsteady Flamelet Modeling of Soot Formation in Turbulent Diffusion Flames

Lecture 9 Laminar Diffusion Flame Configurations

A validation study of the flamelet approach s ability to predict flame structure when fluid mechanics are fully resolved

Best Practice Guidelines for Combustion Modeling. Raphael David A. Bacchi, ESSS

A comparison between two different Flamelet reduced order manifolds for non-premixed turbulent flames

Towards regime identification and appropriate chemistry tabulation for computation of autoigniting turbulent reacting flows

Droplet Behavior within an LPP Ambiance

Overview of Turbulent Reacting Flows

Overview of Reacting Flow

An Unsteady/Flamelet Progress Variable Method for LES of Nonpremixed Turbulent Combustion

Reacting Flow Modeling in STAR-CCM+ Rajesh Rawat

Characterizing the regimes of spray flame-vortex interactions: A spray spectral diagram for extinction

Fluid Dynamics and Balance Equations for Reacting Flows

LES Approaches to Combustion

The effect of momentum flux ratio and turbulence model on the numerical prediction of atomization characteristics of air assisted liquid jets

Modeling of Wall Heat Transfer and Flame/Wall Interaction A Flamelet Model with Heat-Loss Effects

S. Kadowaki, S.H. Kim AND H. Pitsch. 1. Motivation and objectives

Development and Validation of a Partially Coupled Soot Model for Turbulent Kerosene Combustion in Industrial Applications

Computer Fluid Dynamics E181107

IMPROVED POLLUTANT PREDICTIONS IN LARGE-EDDY SIMULATIONS OF TURBULENT NON-PREMIXED COMBUSTION BY CONSIDERING SCALAR DISSIPATION RATE FLUCTUATIONS

Numerical Investigation of Ignition Delay in Methane-Air Mixtures using Conditional Moment Closure

Thermal NO Predictions in Glass Furnaces: A Subgrid Scale Validation Study

LES of an auto-igniting C 2 H 4 flame DNS

DNS of Reacting H 2 /Air Laminar Vortex Rings

Investigation of ignition dynamics in a H2/air mixing layer with an embedded vortex

Combustion. Indian Institute of Science Bangalore

Introduction Flares: safe burning of waste hydrocarbons Oilfields, refinery, LNG Pollutants: NO x, CO 2, CO, unburned hydrocarbons, greenhouse gases G

Numerical Modelling of Soot Formation in Laminar Axisymmetric Ethylene-Air Coflow Flames at Atmospheric and Elevated Pressures

A G-equation formulation for large-eddy simulation of premixed turbulent combustion

Modeling of dispersed phase by Lagrangian approach in Fluent

Well Stirred Reactor Stabilization of flames

Investigation of ignition dynamics in a H2/air mixing layer with an embedded vortex

LARGE-EDDY SIMULATION OF PARTIALLY PREMIXED TURBULENT COMBUSTION

three dimensional Direct Numerical Simulation of Soot Formation and Transport in a Temporally-Evolving Nonpremixed Ethylene Jet Flame

DARS overview, IISc Bangalore 18/03/2014

HIGH-FIDELITY MODELS FOR COAL COMBUSTION: TOWARD HIGH-TEMPERATURE OXY-COAL FOR DIRECT POWER EXTRACTION

The Effect of Flame Structure on Soot Formation and Transport in Turbulent Nonpremixed Flames Using Direct Numerical Simulation

Using Representative Interactive Flamelets in Three- Dimensional Modeling of the Diesel Combustion Process Including Effects of Heat Transfer

TOPICAL PROBLEMS OF FLUID MECHANICS 97

Flame / wall interaction and maximum wall heat fluxes in diffusion burners

Simulation of Turbulent Lifted Flames and their Transient Propagation

A-Priori Analysis of Conditional Moment Closure Modeling of a Temporal Ethylene Jet Flame with Soot Formation Using Direct Numerical Simulation

UNIVERSITA DEGLI STUDI DI PADOVA. High performance computing of turbulent evaporating sprays

Efficient Engine CFD with Detailed Chemistry

ANSYS Advanced Solutions for Gas Turbine Combustion. Gilles Eggenspieler 2011 ANSYS, Inc.

Direct Numerical Simulation of Nonpremixed Flame Extinction by Water Spray

Current progress in DARS model development for CFD

Department of Mechanical Engineering BM 7103 FUELS AND COMBUSTION QUESTION BANK UNIT-1-FUELS

Combustion and Emission Modeling in CONVERGE with LOGE models

Detailed Numerical Simulation of Liquid Jet in Cross Flow Atomization: Impact of Nozzle Geometry and Boundary Condition

MULTIPHASE FLOW MODELLING

Simulation of a lean direct injection combustor for the next high speed civil transport (HSCT) vehicle combustion systems

Turbulent Premixed Combustion

TURBINE BURNERS: Engine Performance Improvements; Mixing, Ignition, and Flame-Holding in High Acceleration Flows

Combustion basics... We are discussing gaseous combustion in a mixture of perfect gases containing N species indexed with k=1 to N:

Large-Eddy Simulation of Reacting Turbulent Flows in Complex Geometries

Towards the prediction of soot in aero-engine combustors with large eddy simulation

A NUMERICAL ANALYSIS OF COMBUSTION PROCESS IN AN AXISYMMETRIC COMBUSTION CHAMBER

NUMERICAL MODELING OF THE GAS-PARTICLE FLUID FLOW AND HEAT TRANSFER IN THE SLIP REGIME

PDF Modeling and Simulation of Premixed Turbulent Combustion

FEDSM-ICNMM

Effects of Damköhler number on flame extinction and reignition in turbulent nonpremixed flames using DNS

Mixing and Combustion in Dense Mixtures by William A. Sirignano and Derek Dunn-Rankin

Construction of Libraries for Non-Premixed Tabulated Chemistry Combustion Models including Non-Adiabatic Behaviour due to Wall Heat Losses

Modeling laminar partially premixed flames with complex diffusion in OpenFOAM R

An Improved Representative Interactive Flamelet Model Accounting for Evaporation Effect in Reaction Space (RIF-ER) SeungHwan Keum

Steady Laminar Flamelet Modeling for turbulent non-premixed Combustion in LES and RANS Simulations

Spray Combustion Modeling Including Detailed Chemistry

Combustion Generated Pollutants

Carbon Science and Technology

Scalar gradient and small-scale structure in turbulent premixed combustion

Topics in Other Lectures Droplet Groups and Array Instability of Injected Liquid Liquid Fuel-Films

Lecture 12. Droplet Combustion Spray Modeling. Moshe Matalon

PDF modeling and simulation of premixed turbulent combustion

Convective Mass Transfer

Reductions of PAH and Soot by Center Air Injection

Triple flame: Inherent asymmetries and pentasectional character

Accounting for spray vaporization in turbulent combustion modeling

Modeling of Humidification in Comsol Multiphysics 4.4

A Zone Model for Fast Verification of Release of Ultrafine Water Mist for Fire Extinction in Compartments

Laminar Premixed Flames: Flame Structure

Fuel, Air, and Combustion Thermodynamics

LES of a Methanol Spray Flame with a Stochastic Sub-grid Model

Theoretical Developments in Group Combustion of Droplets and Sprays

HOT PARTICLE IGNITION OF METHANE FLAMES

NANOPARTICLE COAGULATION AND DISPERSION IN A TURBULENT PLANAR JET WITH CONSTRAINTS

COUNTERFLOW COMBUSTION MODELING OF CH 4 /AIR AND CH 4 /O 2 INCLUDING DETAILED CHEMISTRY

MCS 7 Chia Laguna, Cagliari, Sardinia, Italy, September 11-15, 2011

Combustion: Flame Theory and Heat Produced. Arthur Anconetani Oscar Castillo Everett Henderson

Rocket Propulsion Prof. K. Ramamurthi Department of Mechanical Engineering Indian Institute of Technology, Madras

Simulation of soot formation and analysis of the "Advanced soot model" parameters in an internal combustion engine

Modeling of jet and pool fires and validation of the fire model in the CFD code FLACS. Natalja Pedersen

Modeling Complex Flows! Direct Numerical Simulations! Computational Fluid Dynamics!

Transcription:

Center for Turbulence Research Proceedings of the Summer Program 2006 325 A numerical simulation of soot formation in spray flames By H. Watanabe, R. Kurose, S. Komori AND H. Pitsch A two-dimensional numerical simulation is applied for spray flames formed in a laminar counterflow, and soot formation behavior is studied in terms of equivalence ratio and radiation in detail. N-decane (C 10 H 22 is used as a liquid spray fuel, and the droplet motion is calculated by the Lagrangian method. A one-step global reaction is employed for the combustion reaction model. A kinetically based soot model with flamelet model is employed to predict soot formation. Radiation is taken into account using the discrete ordinate method. The results show that the soot is formed in the spray diffusion flame region and its radiation emission increases with an increase in the equivalence ratio of the droplet fuel. This trend is in good agreement with that of the luminous flame behavior observed in the experiment. The radiation is found to strongly affect the soot formation behavior. Without the radiation model, the soot volume fraction is fatally over-predicted. 1. Introduction Spray combustion is utilized in a number of engineering applications such as energy conversion and propulsion devices. It is, therefore, necessary to precisely predict the spray combustion behavior in designing and operating the equipment such as a gas turbine combustor and diesel engine. However, since the spray combustion is a complex phenomenon in which dispersion of the liquid fuel droplets, their evaporation, chemical reaction of the fuel vapor with oxidizer, etc. take place interactively and simultaneously, the underlying physics governing these processes have not been well understood. Kurose and colleagues (Kurose et al. 2004; Nakamura et al. 2005; Watanabe et al. 2006 have performed numerical simulations of spray flames formed in a laminar counterflow, and investigated the behavior of the spray flames in terms of droplet group combustion and flamelet modeling in detail. In Nakamura et al. (2005, it was mentioned that the behavior of diffusion flame originating from the droplet group combustion is similar to that of the luminous flame observed in the experiment (Hwang et al. 2000 because soot is mainly formed in the diffusion flame. However, since no soot formation model was employed in their computations, the exact soot formation mechanism was not discussed. In recent years, several studies on soot formation using a kinetically based soot model with flamelet model have been reported. The flamelet model is used to give the concentrations of precursors of soot, which are minor chemical species and therefore obtained by solving a number of equations with many chemical species. It is believed that acetylene (C 2 H 2 or polycyclic aromatic hydrocarbons (PAH can be the precursors. Pitsch et al. (2000 investigated the soot formation in a C 2 H 4 jet diffusion flame using a PAH inception model. Wen et al. (2003 evaluated the model dependence on the soot formation between the C 2 H 2 and PAH inception models in their kerosene jet flame simulation. Energy Engineering Research Laboratory, Central Research Institute of Electric Power Industry Department of Mechanical Engineering and Science, Kyoto University

326 H. Watanabe et al. Figure 1. Schematic drawing of the computational setup. These studies show the applicability of the flamelet model to predict the soot formation in gaseous diffusion flames. The purpose of this study is to perform a numerical simulation of spray flames with soot formation, and investigate the effects of equivalence ratio of droplet fuel and radiation on the soot formation. A two-dimensional numerical simulation of spray flames formed in a laminar counterflow is examined. A one-step global reaction is employed for the combustion model of the liquid fuel (n-decane, C 10 H 22. The flamelet model is employed to determine the concentration of the precursor of soot. The radiative heat transfer is calculated using the discrete ordinate method with S 8 quadrature set. 2. Computational setup and numerical methods 2.1. Computational setup The computational setup for spray flames in a laminar counterflow is designed to match the experiment by Hwang et al. (2000. The computational domain considered in this study is shown in Fig. 1. The dimensions of the computational domain normalized by the diameter of the burner ports L p, which are located on both the upper and lower sides, are 1 and 2 in the x and y directions, respectively. The origin of the computational domain is located at the center of the upper port. N-decane (C 10 H 22 is used as the liquid fuel. From the upper port, atmospheric air (T = 300 K, P = 0.1013 MPa, and oxygen mass fraction Y O2 = 0.2357 is issued at 0.5 y 0.5, and pure n-decane spray is injected at 0.15 y 0.15. From the lower port, premixed atmospheric air and vaporized n-decane is issued at 0.5 y 0.5 to stabilize the spray flame. The fluid velocity issued from the upper and lower ports are the same. For the combustion reaction model, a one-step global reaction model (Westbrook & Dryer 1984, is adopted. C 10 H 22 + 31 2 O 2 10CO 2 + 11H 2 O, (2.1 2.2. Governing equations In this study, a kinetically based soot model is employed to predict the soot formation. As a precursor and oxidizers of the soot, acetylene (C 2 H 2 and OH radical and O 2 are

A numerical simulation of soot formation in spray flames 327 chosen, respectively. The concentrations of C 2 H 2 and OH radical are determined using the steady flamelet model, as described later. The governing equations directly solved for the gaseous phase are mass, momentum, energy, and mixture fraction conservation given as ρ t + ρu j = S m, (2.2 ρu i + (ρu i u j + P δ ij σ ij ρg i = S ui, (2.3 t [ ] ρh t + ρu j h ρλ h n + h k (ρλ ρd k Y k = S h, (2.4 ρy k t k=1 + (ρu j Y k ρd k Y k = S combu,k + S Yk, (2.5 ρz t + (ρu j Z ρd Z Z = S Z, (2.6 where u i is the gaseous phase velocity, ρ is the density, P is the static pressure, σ is the stress tensor, g i is the gravitational force, h is the specific total enthalpy, λ is the gaseous thermal diffusivity, and h k, Y k, and D k are the specific enthalpy, the mass fraction, and the mass diffusion coefficient of the k-th species, respectively. δ ij is the Kronecker delta function. Z is the mixture fraction, which is introduced for the flamelet model. The diffusion coefficient of Z, D Z is given by diffusion coefficient of the mixture. For Y k, five major chemical species (O 2, N 2, CO 2, H 2 O, and C 10 H 22 are solved. The gas phase density, ρ is calculated from the equation of state for an ideal gas. The source terms S i due to interactions between gaseous and disperse phase are expressed using the total number of droplets N d existing in the control volume of the gaseous phase calculations, S h = 1 V S ui = 1 V N [ d 1 2 S Yk = S m = 1 V N [ d m d f 1 τ d N d dmd, (2.7 (u i u d,i + dm d u d,i ], (2.8 d (m du d,i u d,i + Q d + dm ] d h V,S + Q rad, (2.9 1 V N d dmd, if k = F (fuel, (2.10 0, otherwise, ( φ 1 N d dmd S Z =, (2.11 φ + Y O2,air V where m d is the droplet mass, u d,i is the droplet velocity, V is the volume of the control volume for the gaseous phase calculation, and h V,S is the evaporated vapor enthalpy at the droplet surface. Q d and Q rad are the heat transfers for the convection and radiation, respectively. Y O2,air is the mass fraction of oxygen in air. The source term S combu,k in

328 H. Watanabe et al. the equations of the species conservation is expressed using the combustion reaction rate per unit volume R F as follows: S combu,k = n k n F W k W F R F, (2.12 where n k and n F are the molar stoichiometric coefficients of the k-th species and the fuel for the one-step global reaction (positive for the production side, respectively. W k and W F are the molecular weights of the k-th species and fuel, respectively. The fuel droplets are tracked individually in a Lagrangian manner. It is assumed that the density of the droplets is much larger than that of the continuous phase such that only the drag and the gravity are significant. The effect of fluid shear on the fluid force acting on the droplets is neglected (Kurose & Komori 1999. Furthermore, droplet breakup, collision, and dense particulate effects are neglected (Ham et al. 2003. Concerning the vaporization of droplets, a non-equilibrium Langmuir-Knudsen evaporation model is chosen (Miller & Bellan 1999. The Lagrangian droplet equations for the position x d,i, velocity u d,i, temperature T d, and mass m d are given by dx d,i = u d,i, (2.13 du d,i = f 1 τ d (u i u d,i + g i, (2.14 = Nu ( cp 3Pr c p,d ( f2 τ d dt d = Q d + (dm d /L V + Q rad m d c p,d (T T d + 1 ( dmd LV + 1 [ Ad ɛ d (πi σtd 4 ], (2.15 m d c p,d m d c p,d dm d = Sh m d ln(1 + B M. (2.16 3Sc τ d Here, T is the gaseous temperature, c p is the specific heat of mixture gas, c p,d is the specific heat of the liquid, L V is the latent heat of vaporization at T d. τ d is the particle response time, and B M is the mass transfer number. f 1 and f 2 are the corrections of the Stokes drag and heat transfer for an evaporating droplet, respectively (Kurose et al. 2003. A d is the projected area of the droplet, ɛ d is the particle emission factor, and σ is the Stefan-Boltzman constant. The radiation intensity I is calculated by the radiative transport equation. The details of the numerical procedure is described in Nakamura et al. (2005 and Watanabe et al. (2006. 2.3. Soot formation modeling The transport equations of the soot number density N and the mass density M are given as ρφ t + (ρu j φ ρd s φ = S φ, (2.17 where D s is the diffusion coefficient of the soot, and φ represents N or M. Here, N and M are given by N = N A ρφ N, M = ρφ M, (2.18

A numerical simulation of soot formation in spray flames 329 respectively. N A is the Avogadro number (6.022 10 26 kmol 1. The source terms for φ N and φ M can be expressed as S N = 1 [( ( ] dn dn +, (2.19 N A Inc. Coa. ( dm ( dm S M = M ( P dn + +, (2.20 N A Inc. Gro. Oxi. is the mass of a soot nucleus, and has a value of 1200 kg/kmol (based on where M P the assumption that the soot size corresponds to 100 carbon atoms. The inception, coagulation, growth, and oxidation rates are calculated in the soot formation model. A simplified soot inception model based on C 2 H 2 concentration (Leung et al. 1991 is used in this study. The inception rate is given by ( dn Inc. = c 1 N A ( ρ Y C 2 H 2 W C2 H 2 e 21100 T, (2.21 where c 1 = 54s 1 (Brookes & Moss 1999. The coagulation of soot particles is assumed to be proportional to the particle collision frequency. The collision frequency is determined by the size of the particles and the mean free path of the surrounding gas. The coagulation rate is given by (Puri et al. 1993 ( ( 1/2 ( 1/6 dn 24R 6 = T 1/2 M 1/6 N 11/6, (2.22 ρ soot N A πρ soot Coa. where R is the universal gas constant, and ρ soot = 2000 kg/m 3. The soot growth model on the surface is based on C 2 H 2 concentration (Frenklach et al. 1984; Harris et al. 1988. The growth rate is given by ( dm Gro. = c 4 ( ρ Y C 2 H 2 W C2 H 2 e 12100 T [ (πn 1/3 ( 6M ρ soot 2/3 ], (2.23 where c 4 = 9000.6 kg m/(kmol s. The OH radical and O 2 are considered as oxidizers in soot oxidation (Neoh et al. 1981; Lee et al. 1962. The soot oxidation rate is given by ( dm Oxi. = c 5 ηρ Y OH W OH T 1/2 (πn 1/3 ( 6M ρ soot 2/3 c 6 ρ Y ( 2/3 O 2 6M e 19778 T T 1/2 (πn 1/3, (2.24 W O2 ρ soot where η = 0.13, c 5 = 105.81 kg m/(kmol K 1/2 s, and c 6 = 8903.51 kg m/(kmol K 1/2 s. The concentrations of C 2 H 2 and OH radical are determined using the steady flamelet model. In this model, the concentrations of C 2 H 2 and OH radical at each position in physical space are identified by examining a flamelet library, which is obtained by solving a one-dimensional flamelet equation in Z space. As parameters to relate the physical space with the Z space, Z and scalar dissipation rate χ, χ = 2D Z Z 2, (2.25 are used. The one-dimensional laminar n-decane/air diffusion flames are calculated in a

330 H. Watanabe et al. counterflow configuration using the commercial software CFX-RIF. A reduced chemical kinetic mechanism with 112 species and 883 elemental reactions is considered for the flamelet library (Ansys, Inc. 2005. In solving the Z equation (2.6, the transport of Z originated from the premixed fuel from the lower port for the stabilization of spray flames is neglected. 2.4. Radiation modeling The radiative heat transfer is based on the discrete ordinate method (Fiveland 1988. The balance of energy passing a specified direction Ω through a small differential volume in an emitting-absorbing medium can be written as follows: Ω I(r, Ω = (α + κi(r, Ω + αi b (r, (2.26 where I(r, Ω is the radiation intensity, which is a function of position and direction, I b (r is the intensity of a blackbody radiation at the temperature of the medium, and α and κ are the absorption and scattering coefficients of the medium, respectively. α is calculated using RADCAL (Grosshandler 1993, a detailed narrow-band model. S 8 quadrature approximation, which corresponds to 80 fluxes, is used to solve for the discrete ordinate directions. The radiative transport equation is iteratively solved with the energy conservation equation (2.4. The heat source by the radiation is given by ( Q rad = α IdΩ σt 4. (2.27 4π 2.5. Computational details The governing equations for the gaseous phase (Eqs. (2.2 (2.6, and (2.17 are solved by a finite volume method using the SIMPLE algorithm. The calculation domain (0 x 1, 1 y 1 is divided into 200 400 equally spaced computational cells in the x and y directions, respectively. The spatial integration is approximated by a fourthorder central difference scheme and the time integration is performed via a fully implicit method. The equations of droplet behavior (Eqs. (2.13 (2.16 are integrated using a second-order Adams-Bashforth scheme. The computations are performed for the strain rate of a = 40 s 1 and the initial droplet size, the Sauter mean diameter (SMD, of 106.7 µm. The initial droplet size distribution is obtained by the Phase Doppler Anemometry (PDA measurement. The strain rate is defined as a ratio of the inlet gaseous velocity u 0 and the distance between the upper and lower port L p (a 2u 0 /L p. To investigate the effect of the equivalence ratio, three computations are performed for φ l = 1.26, 0.84, and 0.63 (corresponding to Cases RE1, RE2, and RE3, respectively with the radiation model. The equivalence ratio is defined as a ratio based on the total mass of the droplets and air issued from the upper port. In addition, to investigate the effect of the radiation on the soot formation, a case without the radiation model (Case E1 with the same a, φ l, and SMD conditions for Case RE1 is computed. Here, four different cases shown in Table 1 are examined in this study. The initial velocity of droplets is the same as the fluid velocity. The Reynolds number based on the burner port diameter (L p = 0.02 m, the fluid velocity (u 0 = 0.4 m/s, and cold air properties is 500. The spray flames stabilized by the planar premixed gas flame issued from the lower port are examined. Computations are run 350 ms (0.1 ms 3, 500 time steps. The CPU time required for each case is approximately 20 hours on AMD Opteron PC.

A numerical simulation of soot formation in spray flames 331 Case Equivalence ratio [-] Strain rate [s 1 ] SMD [µm] Radiation RE1 1.26 40 106.7 Yes RE2 0.84 40 106.7 Yes RE3 0.63 40 106.7 Yes E1 1.26 40 106.7 No Table 1. Cases presented. 3. Results and discussion 3.1. General features Figure 2 shows the instantaneous distributions of (a the gaseous temperature T, (b the soot volume fraction V s, and (c the flame index F I, for Case RE1. White dots indicate the location of the droplets. The parameter F I is used to distinguish between premixed and diffusion flames, and defined as (Yamashita et al. 1996 F I = Y C10 H 22 Y O2. (3.1 F I becomes positive for premixed flames and negative for diffusion flames. The high temperature region formed by burning the fuel from the liquid phase is observed in the center of the flow field in Fig. 2(a. The number of droplets coming from the upper port rapidly decreases in the high temperature region. This indicates that the droplets are vaporized mainly in this region. It is found that the high soot volume fraction is formed in the upper part of the high temperature region (Fig. 2(b. The high soot volume fraction region also corresponds to the location where the diffusion flame originating from droplet group combustion is formed (Fig. 2(c (Nakamura et al. 2005. Figure 3 shows the time averaged profiles of (a the gaseous temperature T, the mixture fraction Z, the soot volume fraction V s, and the evaporation rate of droplets S m, (b the mass fractions of the soot precursor and oxidizers Ȳk, and (c the source terms of the soot mass density S M, respectively. It is found in Fig. 3(a that first, there are main preheat and evaporation layers at 0.25 x/l p 0.3. Then the fuel droplets ignite and T increases to its maximum value. As T increases, Sm increases and reaches its peak value. Since Z produced by the evaporation is transported downstream (in the direction of larger x/l p and keeps growing by droplet evaporation, the peak of Z is located downstream of the peak of S m in the region where the evaporation rate goes to zero. However, the location showing the peak value of V s corresponds to that of Z, and the shape of the profile of V s is similar to that of Z. The mass fraction of OH radical, ȲOH, has two peaks around at x/l p = 0.35 and 0.5, and the peak values of the mass fraction of C 2 H 2, ȲC 2 H 2, and the total soot formation rate appear between the two peaks of ȲOH (Figs. 3(b and (c. The oxidation rate has a negative peak value upstream of the peak of ȲC 2 H 2, and is almost zero downstream. This is because in this downstream region, oxygen has been already consumed by droplet group combustion. 3.2. Effect of equivalence ratio To investigate the effects of equivalence ratio of the liquid phase fuel, the computations are compared among three cases of φ l = 1.26, 0.84, and 0.63 (corresponding to Cases RE1, RE2, and RE3, respectively. Figure 4 shows the comparison of the instantaneous distributions of (a the gaseous temperature T, (b the flame index F I, (c the soot radiation emission E s, and (d photographs taken in the experiment (Hwang et al. 2000.

332 H. Watanabe et al. Figure 2. Instantaneous distributions of (a gaseous temperature T, (b soot volume fraction V s, and (c flame index F I, for Case RE1. Figure 3. Time averaged profiles of (a gaseous temperature T, mixture fraction Z, soot volume fraction V s and evaporation rate of droplets S m, (b mass fractions of soot precursor and oxidizers Ȳ k, and (c source terms of soot mass density S M, for Case RE1.

A numerical simulation of soot formation in spray flames 333 Figure 4. Instantaneous distributions of (a gaseous temperature T, (b flame index F I, (c soot radiation emission E s, and (d experiment s photographs, for φ l = 1.26, 0.84, and 0.63 (Cases RE1, RE2, and RE3, respectively. E s, defined as E s = V s σt 4, (3.2 can be compared with the luminous flame since the brightness of the luminous flame is proportional to the radiation emission from the soot. As φ l increases, both the high gaseous temperature and diffusion flame regions become large (Figs. 4(a and (b. It is also found that E s significantly increases with increasing φ l (Figs. 4(c. This trend is in good agreement qualitatively with that of the luminous flames observed in the experiment (Figs. 4(d. Actually, in the experiment, the luminous flames for Case RE1 are observed constantly, while those for Case RE3 are observed intermittently. Figure 5 shows the time averaged profiles of (a the gaseous temperature T and the evaporation rate of droplets S m, and (b the soot volume fractions V s and mass fractions of C 2 H 2 Ȳ C2 H 2, respectively. Only a small difference in the peak values of T can be observed among Cases RE1, RE2, and RE3 (Fig. 5(a. However, V s increases dramatically as φ l increases (Fig. 5(b. This is due to the increase of ȲC 2 H 2, namely the production of the precursor of the soot becomes marked with increasing S m. 3.3. Effect of radiation To investigate the effect of the radiation, the two cases with and without the radiation (Cases RE1 and E1, respectively are compared. Figure 6 shows comparisons of time averaged profiles of (a the gaseous temperature T and the evaporation rates of droplets

334 H. Watanabe et al. Figure 5. Time averaged profiles of (a gaseous temperature T and evaporation rate of droplets S m, and (b soot volume fraction V s and mass fraction of C 2 H 2 Ȳ C2 H 2, for φ l = 1.26, 0.84, and 0.63 (Cases RE1, RE2, and RE3, respectively. Figure 6. Comparison of time averaged profiles of (a gaseous temperature T and evaporation rates of droplets S m, and (b soot volume fraction V s and mass fraction of C 2 H 2 Ȳ C2 H 2, between the cases of with and without radiation (Cases RE1 and E1, respectively. S m, and (b the soot volume fraction V s and the mass fraction of C 2 H 2 Ȳ C2 H 2. The peak value of T is approximately 300 K higher and the fuel droplets ignite earlier for Case E1 than that for Case RE1 (Fig. 6(a. Also for Case E1, the increase of S m takes place earlier than that for Case RE1. This is because the radiative heat losses, which are caused by the fact that unburned gas and fuel droplets have much lower temperature than the flame, are neglected for Case E1. It is found in Fig. 6(b that V s for Case E1 is much higher than that for Case RE1. This is because T for Case E1 is higher than that for Case RE1, as mentioned earlier. According to Eqs. (2.21 (2.24, like for the soot inception and growth, the oxidation rate should increase with increasing gaseous temperature, and therefore suppress the soot formation. However, for the present cases, oxygen has been already consumed by the droplet group combustion, for which the oxidation is not very effective. As a result, the soot formation is increased by neglecting radiation. 4. Conclusions Two-dimensional numerical simulations were applied for spray flames formed in a laminar counterflow, and the soot formation behavior was studied. The soot is formed in the diffusion flame (droplet group combustion region and its radiation emission increased with an increase in the equivalence ratio of the droplet fuel.

A numerical simulation of soot formation in spray flames 335 This trend is in good agreement with that of the luminous flame behavior observed in the experiment. The increase of the soot radiation emission is found to be due to the increase of the precursor (C 2 H 2 production. It is also found that the radiation strongly affects the soot formation behavior. The soot volume fraction predicted without the radiation model is much higher than that with the radiation model. This is because the temperature of unburned gas and fuel droplets is much lower than the gaseous flame, which decreases the flame temperature and then the soot inception and growth rates through radiation. Acknowledgments The authors are grateful to Prof. Fumiteru Akamatsu of Osaka University for providing the experimental data and useful discussions. The authors would also like to thank Drs. Yuya Baba, Seung-Min Hwang, and Maromu Otaka of CRIEPI, and Dr. Mariko Nakamura of Kobelco Eco-Solutions Co., Ltd. for their help in developing the simulation code. HW is grateful to Dr. Matthias Ihme and Mr. Olivier Desjardins of the Center for Turbulence Research for many inspiring discussions and useful comments on this manuscript. REFERENCES Ansys, Inc. 2005 CFX-RIF User s manual, Ansys, Inc. Brookes, S. J. & Moss, J. B. 1999 Predictions of soot and thermal radiation properties in confined turbulent jet diffusion flames. Combust. Flame 116, 486 503. Fiveland, W. A. 1988 Thee-dimensional radiative heat-transfer solutions by the discrete-ordinates method. J. Thermophysics 2, 309 316. Frenklach, M., Clary, D. W., Gardiner, J., & Stein, S. E. 1984 Detailed kinetic modeling of soot formation in shock-tube pyrolysis of acetylene. Proc. Twentieth Symp. (Int. on Combust., The Combustion Institute, 887 901. Grosshandler, W. L. 1993 RADCAL: A narrow-band model for radiation calculation in a combustion environment. NIST Technical Note 1402. Ham, F., Apte, S. V., Iaccarino, G., Wu, X., Herrmann, M., Constantinescu, G., Mahesh, K. & Moin, P. 2003 Unstructured LES of reacting multiphase flows in realistic gas turbine combustors. CTR Annual Research Brief-2003, Center for Turbulence Research, NASA Ames/Stanford Univ., 135 159. Harris, S. J., Weiner, A. M. & Blint, R. J. 1988 Formation of small aromatic molecules in a sooting ethylene flame. Combust. Flame 72, 91 109. Hwang, S.-M., Saito, H., Takada, S., Akamatsu, F. & Katsuki, M 2000 Observation of spray flame stabilized in a laminar counterflow field. Proc. thirty-eighth Japanese Combust. Symp., 201 202. Kurose, R. & Komori, S. 1999 Drag and lift forces on a rotating sphere in a linear shear flow. J. Fluid Mech. 384, 183 206. Kurose, R., Makino, H., Komori, S., Nakamura, M., Akamatsu, F. & Katsuki, M. 2003 Effects of outflow from the surface of a sphere on drag, shear lift, and scalar diffusion. Phys. Fluids 15, 2338 2351. Kurose, R., Desjardins, O., Nakamura, M., Akamatsu, F. & Pitsch, H. 2004 Numerical simulations of spray flames. CTR Annual Research Briefs-2004, Center for Turbulence Research, NASA Ames/Stanford Univ., 269 280.

336 H. Watanabe et al. Lee, K. B., Thring, M. W. & Beer, J. M. 1962 On the rate of combustion of soot in a laminar soot flame. Combust. Flame 6, 137 145. Leung, K. M., Lindste, R. P. & Jones, W. P. 1991 Simplified reaction mechanism for soot formation in nonpremixed flames. Combust. Flame 87, 289 305. Miller, R. S., Bellan, J. 1999 Direct numerical simulation of a confined threedimensional gas mixing layer with one evaporating hydrocarbon-droplet-laden stream. J. Fluid Mech. 384, 293 338. Nakamura, M., Akamatsu, F., Kurose, R. & Katsuki, M. 2005 Combustion mechanism of liquid fuel spray entering gaseous flame front. Phys. Fluids 17, 123301. Neoh, K. G., Howard, J. B. & Sarofim, A. F. 1981 Particulate Carbon Formation During Combustion, Plenum, 261. Pitsch, H., Riesmeier, E. & Peters, N. 2000 Unsteady flamelet modeling of soot formation in turbulent diffusion flames. Combust. Sci. and Tech. 158, 389 406. Puri, R., Richardson, T. F., Santoro, R. J. & Dobbins, R. A. 1993 Aerosol dynamic processes of soot aggregates in a laminar ethene diffusion flame. Combust. Flame 92, 320 333. Watanabe, H., Kurose, R., Hwang, S.-M. & Akamatsu, F. 2006 Characteristics of flamelet in spray flames formed in a laminar counterflow. Combust. Flame, in press. Wen, Z., Yun, S., Thomson, M. J. & Lightstone, M. F. 2003 Modeling soot formation in turbulent kerosene/air jet diffusion flame. Combust. Flame 135, 323 340. Westbrook, C. K. & Dryer, F. L. 1984 Chemical kinetic modeling of hydrocarbon combustion. Prog. Energy Combust. Sci. 10, 1 57. Yamashita, H., Shimada, M. & Takeno, T. 1996 A numerical study on flame stability at transition point of jet diffusion flames. Proc. Twenty-fifth Symp. (Int. on Combust., The Combustion Institute, 27 34.