Radiation of noise in turbulent non-premixed flames

Similar documents
LES of the Sandia Flame D Using an FPV Combustion Model

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

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

Analysis of different sound source formulations to simulate combustion generated noise using a hybrid LES/APE-RF method

Heating effects on the structure of noise sources of high-speed jets

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

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

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

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

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

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

K. A. Kemenov, H. Wang, S. B. Pope Sibley School of Mechanical and Aerospace Engineering, Cornell University, Ithaca, NY 14853

Application of Compact Schemes to Large Eddy Simulation of Turbulent Jets

The influence of C ϕ is examined by performing calculations with the values C ϕ =1.2, 1.5, 2.0 and 3.0 for different chemistry mechanisms.

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

The development of advanced combustion systems is mainly controlled by the objective to increase fuel

Capturing localised extinction in Sandia Flame F with LES-CMC

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

Subgrid-scale mixing of mixture fraction, temperature, and species mass fractions in turbulent partially premixed flames

LARGE-EDDY SIMULATION OF PARTIALLY PREMIXED TURBULENT COMBUSTION

An evaluation of LES for jet noise prediction

Direct numerical simulation of a turbulent reacting jet

Large Eddy Simulation of Piloted Turbulent Premixed Flame

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

Large-eddy simulation of a bluff-body-stabilized non-premixed flame using a recursive filter-refinement procedure

An evaluation of a conservative fourth order DNS code in turbulent channel flow

Ray traces through unsteady jet turbulence

RANS-Based Noise Predictions of Jets with Internal Forced Mixers *

Large Eddy Simulation of Three-Stream Jets

3-D Large Eddy Simulation for Jet Aeroacoustics

Impact of numerical method on auto-ignition in a temporally evolving mixing layer at various initial conditions

[N175] Development of Combined CAA-CFD Algorithm for the Efficient Simulation of Aerodynamic Noise Generation and Propagation

IMPLEMENTING THE FFOWCS WILLIAMS AND HAWKINGS ACOUSTIC ANALOGY IN ANTARES

Simulation of Turbulent Lifted Flames and their Transient Propagation

Lecture 9 Laminar Diffusion Flame Configurations

Numerical Investigation of the Transonic Base Flow of A Generic Rocket Configuration

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

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

DNS, LES, and wall-modeled LES of separating flow over periodic hills

(A Primitive Approach to) Vortex Sound Theory and Application to Vortex Leapfrogging

Jet Noise Analysis of a Mixed Turbofan Engine

Modeling flame brush thickness in premixed turbulent combustion

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

Semi-Empirical Prediction of Noise from Non-Zero Pressure Gradient Turbulent Boundary Layers

Beyond the Point Particle: LES-Style Filtering of Finite-Sized Particles

PDF modeling and simulation of premixed turbulent combustion

FEDSM COMPUTATIONAL AEROACOUSTIC ANALYSIS OF OVEREXPANDED SUPERSONIC JET IMPINGEMENT ON A FLAT PLATE WITH/WITHOUT HOLE

A Novel FEM Method for Predicting Thermoacoustic Combustion Instability

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

PDF Modeling and Simulation of Premixed Turbulent Combustion

Overview of Turbulent Reacting Flows

Numerical Modeling of Anode Baking Process with COMSOL Multiphysics

Lecture 8 Laminar Diffusion Flames: Diffusion Flamelet Theory

Lecture 14. Turbulent Combustion. We know what a turbulent flow is, when we see it! it is characterized by disorder, vorticity and mixing.

Flamelet Analysis of Turbulent Combustion

Turbulent Premixed Combustion

Before we consider two canonical turbulent flows we need a general description of turbulence.

Large eddy simulation/probability density function modeling of a turbulent CH 4 =H 2 =N 2 jet flame

Post-processing of large-eddy simulations for jet noise predictions

Insights into Model Assumptions and Road to Model Validation for Turbulent Combustion

Fluid Dynamics and Balance Equations for Reacting Flows

Large-Eddy Simulation of Turbulent Combustion

Advanced Turbulence Models for Emission Modeling in Gas Combustion

LES Approaches to Combustion

XXXVIII Meeting of the Italian Section of the Combustion Institute

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

NUMERICAL ANALYSIS OF TURBULENT FLAME IN AN ENCLOSED CHAMBER

Journal of Computational Physics 157, (2000) doi: /jcph , available online at

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

EVALUATION OF FOUR TURBULENCE MODELS IN THE INTERACTION OF MULTI BURNERS SWIRLING FLOWS

CHARACTERISTICS OF ELLIPTIC CO-AXIAL JETS

Examination of the effect of differential molecular diffusion in DNS of turbulent non-premixed flames

Numerical Simulation of Entropy Generation in Hydrogen Enriched Swirl Stabilized Combustion

Prediction of unsteady heat transfer from a cylinder in crossflow

Numerical simulations of the edge tone

VERTICAL TURBULENT BUOYANT HELIUM JET CFD MODELING AND VALIDATION

The JHU Turbulence Databases (JHTDB)

A NUMERICAL ANALYSIS OF COMBUSTION PROCESS IN AN AXISYMMETRIC COMBUSTION CHAMBER

Consistent turbulence modeling in a hybrid LES/RANS PDF method for non-premixed flames

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

Correction: Sub-grid Scale Modeling of Turbulent Spray Flame using Regularized Deconvolution Method

Large eddy simulation of turbulent flow over a backward-facing step: effect of inflow conditions

CFD Simulation of Internal Flowfield of Dual-mode Scramjet

Aero Acoustics Analysis of Car s ORVM

Parabolized Stability Analysis of Jets Issuing from Serrated Nozzles

Model-Based Imaging of Infrared Radiation Intensity from a Turbulent Non-Premixed Jet Flame and Plume

Turbulence Resolution Scale Dependence in Large-Eddy Simulations of a Jet Flame

ON THE SENSITIVITY OF A FREE ANNULAR SWIRLING JET TO THE LEVEL OF SWIRL AND A PILOT JET

Thermoacoustic Instabilities Research

Comparison of direct and indirect combustion noise mechanisms in a model combustor

Computational Aeroacoustics

A Solution Method for the Reynolds-Averaged Navier-Stokes Equation

Investigation of Noise Sources in Turbulent Hot Jets using Large Eddy Simulation Data

Mixing Models for Large-Eddy Simulation of Nonpremixed Turbulent Combustion

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

The JHU Turbulence Databases (JHTDB)

8th AIAA/CEAS Aeroacoustics Conference June 17 19, 2002 Breckenridge, CO

Hybrid CFD/FEM calculations for the aeroacoustic noise radiated from a radial fan

Aeroacoustic Evaluation of an Axial Fan using CFD Methods Frederik Folke, Martin Hildenbrand (ITB Ingenieure GmbH)

Computational Investigations of High-Speed Dual-Stream Jets

Transcription:

Available online at www.sciencedirect.com Proceedings of the Combustion Institute 32 (29) 545 553 Proceedings of the Combustion Institute www.elsevier.com/locate/proci Radiation of noise in turbulent non-premixed flames Matthias Ihme a, *, Heinz Pitsch b, Daniel Bodony c a Department of Aerospace Engineering, University of Michigan, 37 FXB Building, 32 Beal Avenue, Ann Arbor, MI 489, USA b Department of Mechanical Engineering, Stanford University, Stanford, CA 9435, USA c Department of Aerospace Engineering, University of Illinois at Urbana-Champaign, Urbana, IL 68, USA Abstract A model for the prediction of combustion-generated noise in non-premixed flames has been developed. This model is based on Lighthill s acoustic analogy and employs the flamelet/progress variable model to express the excess density as function of mixture fraction and reaction progress variable. In this model, three major sources of sound have been identified, and their individual contribution to the acoustic spectra and overall sound pressure level are analyzed for a nitrogen-diluted methane hydrogen/air flame. The hybrid approach, combining a large-eddy simulation and a computational aeroacoustic method, introduces spurious noise which can pollute the acoustic results. All relevant sources of spurious noise are analyzed, and a physics-based low-pass filter is proposed which eliminates spurious noise due to the convection of acoustic sources. The numerical predictions for both statistical flow field quantities and acoustic results have been validated with experimental data. The good agreement between experiments and simulation highlights the potential of the method for applications to more complex flow configurations and to provide further understanding of combustion noise mechanisms. Ó 29 The Combustion Institute. Published by Elsevier Inc. All rights reserved. Keywords: Aeroacoustics; Combustion-noise; Large-eddy simulation; Computational aeroacoustics; Turbulent non-premixed combustion. Introduction Recent studies project an increase in the world energy consumption by more than 6 percent over the next 3 years []. Over this period, combustion of fossil fuels are predicted to remain a major technology for the energy supply in all end-user sectors. However, accompanied with combustion are unwanted effects such as the emission of pollutants and noise. Particularly the reduction of the noise footprint of aircraft is currently of great * Corresponding author. Fax: + 734 763 578. E-mail address: mihme@umich.edu (M. Ihme). interest in order to meet future stringent noise regulations. Numerical simulation techniques promise to be a viable tool to gain a better understanding of the underlying noise-generating mechanisms and to assist in developing quieter combustion systems. The generation of aerodynamic sound, however, is a rather inefficient process, since only a minute fraction of the thermochemical and kinetic energy of the turbulent reactive flow is radiated as sound. For the prediction of the emitted noise spectra, it is therefore essential to use accurate numerical methods and computational schemes characterized by low dispersion and dissipation. Furthermore, the accurate modeling of coherent structures and other acoustic sources in a turbulent reacting flow 54-7489/$ - see front matter Ó 29 The Combustion Institute. Published by Elsevier Inc. All rights reserved. doi:.6/j.proci.28.6.37

546 M. Ihme et al. / Proceedings of the Combustion Institute 32 (29) 545 553 is essential, since these structures are the main contributors to the radiated noise. The large-eddy simulation (LES) technique has the potential to provide further insight into the noise-generating mechanisms of turbulent reacting flows. In LES, the larger energy-containing scales are time-accurately resolved and only the effect of the smaller and more homogeneous scales requires modeling. The success of LES in the prediction of non-premixed turbulent flames is attributed to the improved description of scalar mixing and the lower complexity of models for the chemical source terms and other unclosed terms. However, the computational cost and the disparity between the characteristic turbulence and acoustic length scales in low-mach number flows restricts the LES application to the near field domain, and acoustic analogies are frequently employed to propagate the acoustic perturbations to the far-field. Lighthill [2] derived an acoustic analogy for the acoustic pressure perturbation, p, by rearranging the Navier Stokes equations to obtain a linear wave operator for a homogeneous, stationary medium on the left hand side, and regarding the right hand side as an equivalent acoustic source term distribution. A model for the prediction of noise from non-premixed flames was developed by Klein & Kok [3]. In this model, infinitely fast chemistry is used and the input parameters for the characterization of the turbulence spectrum are obtained from a steady RANS calculation. A similar model was recently employed for the prediction of noise emissions from premixed swirling flames [4], and good agreement between simulation results and experimental data was obtained. In this respect, it is interesting to point out that acoustic spectra are rather insensitive to changes in geometry and operating conditions for non-premixed and premixed flames [5,6]. However, the spectral characteristics, total sound power, and scaling rules exhibit considerable differences for both combustion modes, and are dependent on the variations in equivalence ratio, fuel composition, and inlet velocity spectrum. Flemming et al. [7] recently combined LES with Lighthill s analogy to predict direct combustion-noise in an open hydrogen flame. In order to partially account for acoustic effects due to the variation of the sound speed, a, inside the flame, they replaced the constant speed of sound in the original formulation by the local sound speed and ignored effects due to the additional term $ y a 2 $ y p, which appears in the inhomogeneous stationary wave equation. The acoustic source term is modeled by the density variation o 2 sq, which was obtained from LES and interpolated from the LES grid to the CAA grid. A different approach is followed by Bui et al. [8,9], in which an acoustic perturbation equation for reacting flows (APE-RF) is used. This formulation accounts for inhomogeneous acoustic media and identifies the unsteady heat release rate as major sound source. Computational results obtained for different jet flame configurations are in good agreement with experimental data. Both approaches employ a multi-linear interpolation to transfer the localized acoustic source term distribution from the fluid domain to the acoustic region. Since the source distribution is localized in a narrow region surrounding the stoichiometric mixture and both computational meshes are typically not aligned in the overlapping region, this procedure introduces errors in the acoustic solution. Additional errors due to the residual scale models and limitations on the numerical resolution of the LES are introduced, and the identification and quantification of these errors is essential for the accurate prediction of the acoustic far-field. Such an error quantification for combustionnoise applications has so far not been presented and the objective of this work is to identify major sources of spurious noise which may result in a miss-prediction of the acoustic far-field spectra. In order to accomplish this, a numerical method for the prediction of combustion-noise is first presented. In this method, a flamelet-based combustion model is employed to extend Lighthill s acoustic analogy to account for effects due to the unsteady reaction rate and density variations on the radiated noise. This homogeneous-media wave equation is solved in Fourier-space for arbitrary far-field locations using an integral formulation. The remainder of this article is organized as follows. The combustion model and the aeroacoustic model are summarized in the next section, and the experimental configuration and numerical setup are presented in Section 3. Statistical flowfield results are discussed in Section 4. Sources of spurious noise and computational results for the acoustic pressure signal, source term intensity, and sound pressure levels are analyzed in Section 5. The paper finishes with conclusions. 2. Mathematical formulation In the present work, Lighthill s acoustic analogy is employed to predict the radiated sound field emitted from a non-premixed turbulent flame. In this analogy, the noise-generating sources in a limited flow domain X F, are represented by an acoustically equivalent source term distribution, embedded in a homogeneous stagnant fluid region X A, having constant reference properties. For the simulation of the turbulent reactive flow-field, a low-mach number variable density

M. Ihme et al. / Proceedings of the Combustion Institute 32 (29) 545 553 547 LES formulation is used. This is motivated by the observation that for a turbulent reactive jet, the exit Mach number, M = U ref /a ref, based on the jet bulk exit velocity and sound speed in air at ambient temperature, is typically very low to ensure flame stability. 2.. Governing equations and combustion model In low-mach number variable density LES, the instantaneous Favre-filtered conservation equations for mass and momentum are solved in the fluid domain X F. The filter operator introduces a residual stress tensor in the governing equations, which is modeled by a dynamic procedure [,]. In addition, the filtered density and all filtered molecular properties are also unclosed quantities, and a state equation is required to close the set of equations. In the following, this state equation is obtained from the flamelet/progress variable (FPV) combustion model, which relates all thermochemical quantities to two scalars, namely the mixture fraction Z and the reaction progress variable C [2,3]. This reactive scalar corresponds to a linear combination of major product mass fractions, and is here defined as C ¼ Y CO2 þ Y CO þ Y H2 O þ Y H2. In this model, all chemical states of a burning flame are obtained from the solution of the steady flamelet equations, and the FPV state relation, in which all thermochemical quantities are parameterized by Z and C, can then be written as w ¼ G w ðz; CÞ, and w denotes the vector of all thermochemical quantities, viz. w ¼ðq; m; a; Y;...; _x C Þ T. Here, the density is denoted by q, m is the kinematic viscosity, a is the molecular diffusivity, Y is the vector of all species mass fractions, and _x C is the chemical reaction rate of the reaction progress variable. For application in LES, the state relation G w must be expressed in terms of Favre-filtered quantities. This is accomplished by employing a presumed PDF approach, in which a beta distribution is used as model for the marginal PDF of the mixture fraction, and a mixture fraction-conditioned Dirac function is used for the progress variable. With this, all Favre-filtered thermochemical quantities, which are denoted by a tilde, can be written as ew ¼ eg w ðez; g Z 2 ; ecþ; ðþ and g Z 2 is the residual mixture faction variance. More details on the combustion model are given in Refs. [2,3] and are here omitted. In addition to the solution of the Navier Stokes equations, the FPV model requires then the solution of the following transport equations for ez and ec qd s ez ¼ ReSc r y ðqear y ezþþr y s res Z ; ð2aþ qd s ec ¼ ReSc r y ðqear y ecþþr y s res C þ Daq e _x C ; ð2bþ and a dynamic procedure is used to model the residual scalar fluxes s res w ¼ qeu w e quw f [2]. The residual mixture fraction variance, appearing in Eq. () is computed from an algebraic closure model [4]. In the above equations, ed s ¼ o s þ eu r y denotes the Favre-filtered substantial derivative, and the time-space coordinate (s,y) refers to a location in X F. Note, that in this paper all quantities are non-dimensionalized with reference flow-field quantities and are written in non-dimensional form [5]. The Reynolds number is denoted by Re, Sc is the Schmidt number, and Da is the Damköhler number [5]. 2.2. Acoustic model In a turbulent flame, sound is generated by both the turbulent flow field and the combustion in X F. The sound propagates as acoustic waves to the point of perception (t,x) inx A. Lighthill [2] derived the following inhomogeneous wave equation for the pressure disturbance p ¼p p ref M 2 o 2 s p D y p ¼r y r y ðqeueuþ o 2 s q e r y r y r rres ; ð3þ Re where r and r res are the viscous and residual stress tensors, respectively, and q e = q M 2 p is the excess density [6]. By employing the FPV state relation, Eq. (), together with Eq. (2), the excess density can be reformulated. Then, after neglecting contributions due to viscous stresses and all residual scale contributions, Eq. (3) can be written as M 2 o 2 s p D y p ¼r y r y ðqeueuþ fflffl{zfflffl} T R o s rðð qþeuþ fflfflfflfflfflffl{zfflfflfflfflfflffl} F M _x Dao s q g! C q o Cq : ð4þ fflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflffl} The term T R represents a quadrupole source due to unsteady Reynolds stresses, and F M is a fluctuating mass flux of dipole nature. The term Q R denotes a monopole due to the chemical reaction rate of the progress variable, and is precomputed and stored in the flamelet library. Note that Eq. (4) is a homogeneous-media wave equation and does not account for refraction effects due to sound speed variations in the source Q R

548 M. Ihme et al. / Proceedings of the Combustion Institute 32 (29) 545 553 region. Potential implications of this model assumption are discussed in Section 5. An integral formulation for the far-field pressure signal as function of frequency x can be obtained by applying a free-space Green s function and a Fourier transformation to Eq. (4), viz. bp ðx; xþ ¼ Z Z Z ½M 2 jðx 2 Þ : T b 4p R X F þ Mkðx 2 ÞF b M DalðxÞQ b R Šdy; ð5þ and j, k, and l are directional cosine factors as result of the partial integration. By employing a free-space Green s function, it is assumed that all acoustic sources radiate undisturbed to the far-field, and effects due to diffraction at the nozzle or other obstacles are negligible. Diffraction effects are typically only relevant for high frequencies [7], and its significance is dependent on the directivity angle. As such, they have usually insignificant effect on radiation in forward direction, but become increasingly important with increasing angle to the jet axis. Nevertheless, in the experiment that will be used here for model validation, such effects were minimized by covering the orifice and tubing with sound-absorbing material [8,9]. The dependence of the three source terms on Mach- and Damköhler numbers is separately indicated in the prefactors of these terms, and the asymptotic dependence of the three coefficients on x for large distances between source and observer locations is shown in parentheses. More details on the derivation of Eq. (5) can be found in Ref. [5]. 3. Experimental configuration and numerical setup The N 2 -diluted CH 4 H 2 /air flame was experimentally studied by Bergmann et al. [2], Meier et al. [2], and Schneider et al. [22]. The burner configuration for the non-premixed flame consists of a central fuel nozzle of diameter D ref = 8 mm, which is surrounded by a co-flow nozzle of square shape. The fuel bulk velocity is U ref = 42.2 m/s. Co-flow air is supplied at an axial velocity of 7. 3 U ref. The jet fluid consists of a mixture of 22. % methane, 33.2 % hydrogen, and 44.7 % nitrogen by volume with a stoichiometric mixture fraction of Z st =.67. Spectral noise emissions and sound pressure levels of this flame have been quantified by Singh et al. [8,9]. The Reynolds number, based on the nozzle diameter, pipe bulk exit velocity, and kinematic viscosity of the fuel mixture, is Re = 4,74. The Schmidt number is Sc =.486, the Mach number is M =.23, and the Damköhler number is Da =.644. The Favre-filtered conservation equations for mass, momentum, mixture fraction, and progress variable are solved in a cylindrical coordinate system in X F with y =(r,u,y) T. The computational domain is 4 2p 2 in radial, circumferential, and axial directions, respectively. The radial direction is discretized by 6 unevenly spaced grid points concentrated in the fuel nozzle. For the discretization of the jet radius, 3 grid points are used. The circumferential direction is equally spaced and uses 64 points. The grid in axial direction uses 32 points and is, beginning at the nozzle exit, stretched in stream-wise direction. The total number of grid points used for the simulation is approximately 3.28 million. The turbulent inlet velocity profile is generated by separately performing a periodic pipe flow simulation by enforcing a constant mass flux. Convective outflow conditions are used at the outlet and slip-free boundary conditions are employed at the radial boundaries. Flow-field statistics are collected over five flowthrough-times after a statistically steady flow-field was obtained. For the computation of acoustic results, 4,96 samples separated by a sample width of Ds =.4 were used. This data set was subdivided into 5 overlapping intervals to increase statistical sampling, which is common practice in signal processing. The Strouhal number is defined as St = fd ref /U ref, with f denoting the frequency. The computed mean and resolved variance are obtained by averaging in temporal and azimuthal directions, and are denoted by hwiðr; e yþ and hw e 2 < iðr; yþ, respectively, for a scalar quantity w. 4. Flow-field results Statistical results for axial velocity and mixture fraction along the centerline of the jet are shown in Fig.. Symbols denote experimental data and lines correspond to the computational results. a b.5.25.75.5.25.8.6.4.2 2 4 6 8 Fig.. Comparison of measured and calculated mean and resolved rms statistics of (a) axial velocity and (b) mixture fraction along the centerline..2.5..5..8.6.4.2

M. Ihme et al. / Proceedings of the Combustion Institute 32 (29) 545 553 549.8.6.4.2.8.6.4.2.8.6.4.2 -.3 -.2 -...2.3.5.4.3.2..5.4.3.2..5 Fig. 2. Comparison of radial profiles for mean and resolved fluctuations of the mixture fraction at three axial locations in the jet flame..4.3.2..25.75.5.25.25 Apart from the slight over-prediction of the decay rate for 6 y 6 3, the prediction of the mean axial velocity, shown in Fig. a, is in good agreement with experimental data in the self-similar region of the jet flame. The predicted velocity fluctuations, shown by the dashed line, are also in good agreement with experimental data. The accurate prediction of the mixture fraction field is crucial for the description of scalar and temperature distributions in the flame. The evolution of the mean and root-mean-square (rms) mixture fraction along the jet centerline is shown in Fig. b. The numerical results for the mean mixture fraction in the rich part of the flame, i.e., for y 6 65, are in good agreement with the measurements. The slight over-prediction of hez i further downstream results in an over-prediction of the temperature by approximately K (not shown) due to the strong temperature sensitivity in the fuel-lean part of the flame. The resolved mixture fraction variance can considered to be in excellent agreement with the experimental data. Radial profiles of the mean mixture fraction and resolved mixture fraction variance are presented in Fig. 2 for three axial locations in the jet flame. Apart from the slight over-prediction of the peak rms mixture fraction in the shear layer for y =, the computational results can considered to be in excellent agreement with measurements. Profiles of the normalized temperature, H, are compared with experimental data in Fig. 3. It can be seen that the heat release close to the nozzle occurs in a narrow region. The width of this zone increases with increasing downstream distance. The rms temperature profiles at the first two measurement stations exhibit two pronounced peaks which are separated by a local minimum. The reason for this can be attributed to the temperature sensitivity with respect to the mixture fraction around the stoichiometric condition, and has been discussed in Ref. [23]. 5. Acoustic results.75.5.25.25.75.5.25 In the following, acoustic results are discussed that are obtained by evaluating Eq. (5). In this equation, the volume integral extends over the entire LES domain up to 2 diameter in the downstream direction. In a sensitivity study it was found that this integral converges for y P 8, which confirms that all major acoustic sources are contained in the LES domain. 5.. Spurious noise sources.5 The conversion of thermal and kinetic energy into acoustic power is typically a rather inefficient process, especially at low-mach numbers. Because of this inefficiency, only a minute fraction of the total energy in a jet is radiated as sound to the far-field. For the accurate prediction of the farfield sound, it is therefore not only important to employ non-dispersive and non-dissipative numerical methods for the sound propagation, but also provide an accurate characterization of the acoustic source term distribution. These sources are embedded in a turbulent reactive flow-field, whose spatial extend can considerably be larger than the source region of isothermal jets.5.4.3.2..4.3.2..5 -.3 -.2 -...2.3 Fig. 3. Comparison of radial profiles for mean and resolved fluctuations of the temperature at three axial locations in the jet flame..4.3.2.

55 M. Ihme et al. / Proceedings of the Combustion Institute 32 (29) 545 553 at comparable operating conditions. The size of the computational domain in LES can therefore be limited by the required numerical resolution. Furthermore, the different physical models employed in LES can introduce errors which may only have a second order effect on the prediction of statistical flow-field quantities; however, they can pollute the aeroacoustic results. In a simulation of turbulent reactive flows, the following sources of spurious noise can be identified:. LES residual model 2. Combustion model 3. Strouhal number limitation of the acoustic samples 4. Strouhal number limitation of the computational grid e-5 e-6 e-7 e-8 e-3 e-4 e-5 e-6. 3 Fig. 4. Temporal energy spectra of the acoustic source term Q R, multiplied by its x-dependence (see Eq. (5)) at two different locations in the jet flame. e-2 e-3 e-4 e-5 e-3 e-4 e-5 e-6.5. 3 Fig. 5. Temporal energy spectra of the density, appearing in the acoustic source term F M, multiplied by its x 2 - dependence (see Eq. (5)) at two different locations in the jet flame. The vertical dashed lines show the critical advection Strouhal number St a. See Fig. 5 for legend. The residual scale models in LES can result in unphysical noise at high frequencies, corresponding to the unresolved turbulent scales. In the present work, a dynamic procedure was employed [2]. In order to assess the effect of the residual scale viscosity and diffusivity on the acoustic spectra, a highly resolved LES of a reduced computational domain was performed, for which the energy spectra were evaluated. In this study, it was concluded that the residual scale model is an insignificant contributor to the high-frequency noise. Another important source of unphysical noise can be the FPV combustion model, which employs a tabulation method for the parameterization of all thermochemical quantities. Pierce & Moin [2] pointed out that the low-mach number variable density formulation can be affected by spurious heat release which is caused by dispersion errors in the convection terms of the equations for species transport and mass conservation. While the effect of this on the high-frequency noise contribution was not quantified in this work, it was found that the accurate evaluation of the state relation greatly reduces the high frequency error. Examples are shown in Figs. 4 and 5, which compare temporal spectra of two source term contributions obtained with increasing table resolutions at different locations in the jet flame. It can be seen that spurious noise at high frequencies is introduced when a poorly resolved chemistry representation with only 25 grid points in ez and ec directions is used. The spurious noise can further be reduced by employing lookup tables with finer resolutions. Spectra that are shown by solid lines in both figures correspond to results in which a tetrahedral integration method [24] was used. In this method, a computational cell is subdivided into tetrahedrons. The state equation is then integrated for each sub-cell using high-order quadrature rules. Employing this method resulted in better convergence and numerical stability, and was therefore used throughout this work. A detailed analysis of this method, which is not presented here, confirmed that a converged solution is obtained using a table resolution with grid points in each ez and ec directions. Note however that depending on the integration scheme, the computational cost of the simulation can increase considerably. The acoustic source terms were collected at a constant sampling time of Ds =.4. From this, the Nyquist Strouhal number can be computed as St N =(2Ds) = 2.5. Note that the flow solver is advanced with a time increment of Ds =.2, corresponding to a Strouhal number of 25. Therefore, the temporal resolution is sufficiently fine to resolve all relevant audible frequencies. The acoustic sources are generated by a turbulent reactive flow-field, which is advected in

M. Ihme et al. / Proceedings of the Combustion Institute 32 (29) 545 553 55 downstream direction. A critical Strouhal number associated with the advection of the acoustic source terms and the mesh spacing can be written as St a ðyþ ¼heuiðyÞ=ð2DyðyÞÞ. Here, it is assumed that the axial velocity component is the dominant advection velocity, which is valid for turbulent jet flows. The decay of the mean axial velocity in the self-similar region of a turbulent round jet is heui /y, and the axial grid stretching in the present LES, is Dy / y. Thus, the critical Strouhal number can be approximated roughly as St a / y 2. For application in LES, the critical advection Strouhal number is evaluated from the computed mean axial velocity and the grid spacing, and is shown in Fig. 5 by the vertical dashed lines. It can be seen that St a correlates well with the onset of the spurious noise in the high-frequency range of the spectra. Since St a represents the limiting frequency for which the acoustic predictions are reliable, a sharp spectral cut-off filter of the form Hð2pSt a ðyþ xþ is introduced, where H denotes the Heaviside function. 5.2. Instantaneous pressure field Using Eq. (5), the pressure signals resulting from each of the three source terms are computed as function of x and x. These pressure fields are then transformed into the time domain and are shown in Fig. 6 for a particular time instant. For reference, the corresponding source term distributions are also shown. Note that only the axial components of T R and F M are shown. It is interesting to point out that all three source terms have a different spatial distribution. The source term Q R is mainly aligned with the location of the stoichiometric mixture and extends up to 5 nozzle diameters in downstream direction. Similarly, F M extends to the stoichiometric flame length. The source term T R is constrained to the shear layer region in the jet flame, and extends to about 5 diameters in the streamwise direction. Note, however, that for subsonic flows only a small portion of the source term components are radiating. An interpretation of the noise radiation and the role of source convection in wavenumber-frequency space has been given by Crighton [25]. The different source term characteristics which were also discussed in Section 2.2 result in different pressure fields. The instantaneous pressure field that is produced by T R is shown in Fig. 6a. This source is mainly radiating in the 45 forward direction. Compared to the other two pressure fields shown in Fig. 6b and c, the characteristic wave length is considerably smaller. Furthermore, the magnitude of the pressure wave emitted by the source T R is compared to that generated by F M and Q R smaller by one and two orders of magnitude, respectively. Compared to the quadrupole source, the directivity of the pressure signal generated by F M (Fig. 6b) is less pronounced. Interestingly, the radiation pattern of the chemical source (Fig. 6c) shows a pressure perturbation of rather large wave length. A detailed analysis of the temporal evolution of the pressure and the source term spectrum suggests that the acoustic source responsible for this low-frequency emission is mostly confined to the spatial location around y = 3 diameter. 5.3. Pressure spectra The computed sound pressure level (SPL) at two far-field locations and the individual source term contributions are presented in Fig. 7. Experimental data, measured by Singh et al. [8,9], are shown by symbols and computational results are shown by lines. This figure shows that the quadrupole source term T R is the least efficient contributor to the total sound pressure level. The reason for this Fig. 6. Instantaneous pressure distribution in the acoustic domain.

552 M. Ihme et al. / Proceedings of the Combustion Institute 32 (29) 545 553 8 6 4 2 8 6 4 2.5. 5 Fig. 7. Acoustic source term contribution to the SPL at two far-field locations. can be deduced from Eq. (5), which shows that this term scales quadratically with the Mach number. The fluctuating mass flux contribution is about 5 db lower than the total SPL in the low- and mid-frequency range, which can be attributed to its linear dependence on the Mach number. The dominant source term contribution to the total SPL is the chemical source term. This suggests that for the prediction of the SPL at low and mid frequencies, the source terms T R and F M can essentially be neglected, and the total SPL radiated from an open non-premixed flame can be sufficiently characterized by the chemical source term alone. It is interesting to point out that some phase cancelation can be observed in the side line direction around St =.. This cancelation is a result of the phase shift between the three different source term contributions, and results in a total SPL which is approximately 2 db lower than the SPL generated by the chemical source term alone. The prediction shows that the SPL is approximately constant for St 6.3, which is in agreement with the experimental data, and also with the theoretical analysis by Strahle [26]. The numerical simulation over-predicts the SPL for St in the 45 forward direction. A careful analysis of the experimental data and further theoretical analysis provides some evidence that this is caused by high-frequency refraction effects which are only insufficiently captured with Lighthill s acoustic analogy and the assumption of constant sound speed [27,28]. In Ref. [27] it was shown that acoustic refraction due to temperature-dependent variations in the sound speed are mostly relevant in the jet forward direction. Highfrequency acoustic waves that propagate in the downstream direction interact with sound speed gradients, and are refracted away from the jet axis. This leads to the formation of a so-called zone of silence, which increases with increasing frequency. 6. Conclusions A model for the prediction of combustion-generated noise in non-premixed turbulent flames has been presented. This model is based on Lighthill s acoustic analogy and employs the FPV assumption in order to express the excess density in terms of mixture fraction and progress variable. The resulting wave equation is converted into an integral formulation and solved in Fourier-space. The following conclusions can be drawn: Mainly three acoustic sources have been identified as contributors to the far-field pressure, and the computed SPL confirms that the chemical source term is the main contributor to the far-field noise for low-mach number flows. The hybrid LES/CAA formulation introduces spurious noise into the prediction of the acoustic pressure. All relevant sources have been analyzed, and it was found that the limitation of the numerical grid resolution and the chemistry representation are the major sources of spurious noise. A physics-based low-pass filter has been proposed, which eliminates spurious noise introduced due to the convection of acoustic sources. The SPL is in good agreement with experimental data; some discrepancies in the SPL in the jet forward direction at high frequencies are apparent and are attributed to refraction effects which are not fully captured with the present model formulation. Acknowledgments The authors gratefully acknowledge funding by the United States Department of Energy within the Advanced Simulation and Computing (ASC) program. Helpful discussions with Sanjiva Lele are gratefully acknowledged. We thank Andreas Dreizler, Jay P. Gore, and Kapil K. Singh for sharing their experimental data with us. References [] International Energy Outlook. DOE/EIA- 484(27), 27. Available at www.eia.doe.gov/ oiaf/ieo/index.html. [2] M.J. Lighthill, Proc. R. Soc. London A 2 (7) (952) 564 587. [3] S.A. Klein, J.B.W. Kok, Combust. Sci. Tech. 49 (999) 267 295.

M. Ihme et al. / Proceedings of the Combustion Institute 32 (29) 545 553 553 [4] C. Hirsch, A. Winkler, J. Wäsle, T. Sattelmayer, in: 3th International Congress on Sound and Vibration, Vienna, Austria. [5] K.K. Singh, C. Zhang, J.P. Gore, L. Mongeau, S.H. Frankel, Proc. Combust. Inst. 3 (24) 77 75. [6] R. Rajaram, T. Lieuwen, Combust. Sci. Tech. 75 (2) (23) 2269 2298. [7] F. Flemming, A. Sadiki, J. Janicka, Proc. Combust. Inst. 3 (27) 389 396. [8] T.P. Bui, M. Meinke, W. Schröder, F. Flemming, A. Sadiki, J. Janicka, in: th AIAA/CEAS Aeroacoustics Conference, paper AIAA-25-34. Monterey, USA, May 25. [9] T.P. Bui, M. Ihme, M. Meinke, W. Schröder, H. Pitsch, in: 3th AIAA/CEAS Aeroacoustics Conference, paper AIAA-27-346. Rome, Italy, 27. [] M. Germano, U. Piomelli, P. Moin, W.H. Cabot, Phys. Fluids A 3 (7) (99) 76 765. [] P. Moin, K. Squires, W. Cabot, S. Lee, Phys. Fluids A 3 () (99) 2746 2757. [2] C.D. Pierce, P. Moin, J. Fluid Mech. 54 (24) 73 97. [3] M. Ihme, C.M. Cha, H. Pitsch, Proc. Combust. Inst. 3 (25) 793 8. [4] C.D. Pierce, P. Moin, Phys. Fluids (2) (998) 34 344. [5] M. Ihme, Pollutant Formation and Noise Emission in Turbulent Non-Premixed Flames, Ph.D. thesis, Stanford University, 27. [6] D.G. Crighton, A.P. Dowling, J.E. Ffowcs Williams, M. Heckl, F.G. Leppington, Modern Methods in Analytical Acoustics: Lecture Notes, Springer, London, 992. [7] W.C. Strahle, Proc. Combust. Inst. 4 (973) 527 535. [8] K.K. Singh, S.H. Frankel, J.P. Gore, AIAA J. 4 (2) (23) 39 32. [9] K.K. Singh, S.H. Frankel, J.P. Gore, AIAA J. 42 (5) (24) 93 936. [2] V. Bergmann, W. Meier, D. Wolff, W. Stricker, Appl. Phys. B 66 (4) (998) 489 52. [2] W. Meier, R.S. Barlow, Y.L. Chen, J.-Y. Chen, Combust. Flame 23 (3) (2) 326 343. [22] C. Schneider, A. Dreizler, J. Janicka, E.P. Hassel, Combust. Flame 35 ( 2) (23) 85 9. [23] M. Ihme, H. Pitsch, Combust. Flame (28), doi:.6/j.combustflame.28.4.5. [24] L. Shunn, F. Ham, in: Annual Research Breifs, Center for Turbulence Research, 26, pp. 35 47. [25] D.G. Crighton, Prog. Aerospace Sci. 6 () (975) 3 96. [26] W.C. Strahle, Prog. Astronaut. Aeronaut. 37 (975) 229 248. [27] M. Ihme, M. Kaltenbacher, H. Pitsch, in: Proceedings of the Summer Program, Center for Turbulence Research, 26, pp. 497 5. [28] M. Kaltenbacher, M. Ihme, H. Pitsch, in: 3th AIAA/CEAS Aeroacoustics Conference, paper AIAA-27-343, Rome, Italy, 27.