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

Similar documents
Characterization of Flow Field Structure and Species Composition in a Shear Coaxial Rocket GH2/GO2 Injector: Modeling of Wall Heat Losses. D. J.

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

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

Lecture 8 Laminar Diffusion Flames: Diffusion Flamelet Theory

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

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

Fluid Dynamics and Balance Equations for Reacting Flows

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

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

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

LARGE-EDDY SIMULATION OF PARTIALLY PREMIXED TURBULENT COMBUSTION

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

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

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

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

LES of the Sandia Flame D Using an FPV Combustion Model

Process Chemistry Toolbox - Mixing

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

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

Capturing localised extinction in Sandia Flame F with LES-CMC

D. VEYNANTE. Introduction à la Combustion Turbulente. Dimanche 30 Mai 2010, 09h00 10h30

Laminar Premixed Flames: Flame Structure

Lecture 9 Laminar Diffusion Flame Configurations

Advanced Turbulence Models for Emission Modeling in Gas Combustion

Lecture 6 Asymptotic Structure for Four-Step Premixed Stoichiometric Methane Flames

Combustion Reaction Model Generation using Principal Component Analysis

Flow and added small-scale topologies in a turbulent premixed flame

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

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

Simulation of Turbulent Lifted Flames and their Transient Propagation

LEWIS NUMBER EFFECTS ON THE STRUCTURE AND EXTINCTION OF DIFFUSION FLAMES DUE TO STRAIN. A. Liñan

Fundamental Concepts of Convection : Flow and Thermal Considerations. Chapter Six and Appendix D Sections 6.1 through 6.8 and D.1 through D.

Problem 4.3. Problem 4.4

Predicting NO Formation with Flamelet Generated Manifolds

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

Laminar Forced Convection Heat Transfer from Two Heated Square Cylinders in a Bingham Plastic Fluid

Lecture 10 Turbulent Combustion: The State of the Art

Heat-recirculating combustors. Paul D. Ronney

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

A NUMERICAL ANALYSIS OF COMBUSTION PROCESS IN AN AXISYMMETRIC COMBUSTION CHAMBER

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

LES Approaches to Combustion

Exercises in Combustion Technology

DNS of Reacting H 2 /Air Laminar Vortex Rings

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

CFD and Kinetic Analysis of Bluff Body Stabilized Flame

Unsteady Flamelet Modeling of Soot Formation in Turbulent Diffusion Flames

Flamelet Analysis of Turbulent Combustion

Direct numerical simulation of a turbulent reacting jet

Simulation of a Shear Coaxial GO2/GH2 Rocket Injector with DES and LES Using Flamelet Models

Overview of Turbulent Reacting Flows

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

DNS STUDY OF TURBULENT HEAT TRANSFER IN A SPANWISE ROTATING SQUARE DUCT

UQ in Reacting Flows

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

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

Effects of turbulence and flame instability on flame front evolution

A Priori Model for the Effective Lewis Numbers in Premixed Turbulent Flames

Well Stirred Reactor Stabilization of flames

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

A Priori Testing of Flamelet Generated Manifolds for Turbulent Partially Premixed Methane/Air Flames

Large-Eddy Simulation of Turbulent Combustion

Turbulent Premixed Combustion

6.2 Governing Equations for Natural Convection

Faculty of Engineering. Contents. Introduction

Analysis of homogeneous combustion in Monolithic structures

Introduction to Heat and Mass Transfer. Week 12

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

CONVECTIVE HEAT TRANSFER

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

CONVECTIVE HEAT TRANSFER

DARS overview, IISc Bangalore 18/03/2014

Large-eddy simulations of turbulent reacting stagnation point flows

TOWARDS AN EXTENSION OF TFC MODEL OF PREMIXED TURBULENT COMBUSTION

Convection. forced convection when the flow is caused by external means, such as by a fan, a pump, or atmospheric winds.

Interactions between oxygen permeation and homogeneous-phase fuel conversion on the sweep side of an ion transport membrane

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

Modelling Detailed-Chemistry Effects on Turbulent Diffusion Flames using a Parallel Solution-Adaptive Scheme

Development of One-Step Chemistry Models for Flame and Ignition Simulation

Assessment of a flame surface density-based subgrid turbulent combustion model for nonpremixed flames of wood pyrolysis gas

6. Laminar and turbulent boundary layers

External Forced Convection :

Asymptotic Structure of Rich Methane-Air Flames

Multi-dimensional transport: DNS analysis and incorporation into the Reaction-Diffusion Manifold (REDIM) method

Forced Convection: Inside Pipe HANNA ILYANI ZULHAIMI

Sensitivity analysis of LES-CMC predictions of piloted jet flames

Application of FGM to DNS of premixed turbulent spherical flames

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

Large-Eddy Simulation of a Jet-in-Hot-Coflow Burner Operating in the Oxygen-Diluted Combustion Regime

DNS and LES of Turbulent Combustion

ADVANCED DES SIMULATIONS OF OXY-GAS BURNER LOCATED INTO MODEL OF REAL MELTING CHAMBER

Numerical Investigation of Thermal Performance in Cross Flow Around Square Array of Circular Cylinders

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

A combined application of the integral wall model and the rough wall rescaling-recycling method

Spray evaporation model sensitivities

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

Algebraic Turbulence-Chemistry Interaction Model

Published in: Proceedings of the Fluent Benelux User Group Meeting, 6-7 October 2005, Wavre, Belgium

Summary of Dimensionless Numbers of Fluid Mechanics and Heat Transfer

Convective Mass Transfer

Transcription:

9 th U. S. National Combustion Meeting Organized by the Central States Section of the Combustion Institute May 17-20, 2015 Cincinnati, Ohio Modeling of Wall Heat Transfer and Flame/Wall Interaction A Flamelet Model with Heat-Loss Effects Hao Wu 1*, Matthias Ihme 1, 1 Department of Mechanical Engineering, Stanford University, Stanford, CA 94305, USA * Corresponding Author Email: wuhao@stanford.edu Abstract: An extension of the flamelet/progress variable model is developed to include wall- heat loss effects due to convective heat- transfer. The model introduces a modified thermal boundary condition to the counter- flow diffusion flame configuration in the composition space. The thermochemical composition of the resulting non- adiabatic flamelet structure forms a three- dimensional manifold, which is parameterized in terms of mixture fraction, a progress variable, and temperature. The performance of the model is evaluated in an a priori study of a H2/O2 diffusion flame that is stabilized at an inert isothermal wall. Comparisons with DNS- data show that the developed non- adiabatic flamelet model accurately represents results for temperature, chemical composition, and wall heat transfer. Keywords: flamelet model, wall heat transfer, flame-wall interaction 1. Introduction Turbulent combustion processes with convective heat-losses and flame-wall interaction have presence in many engineering applications. Specifically, they are directly related to the heat transfer and engine cooling systems for rocket engines, to which many of the main-engine failures during the early design and development phase of the space shuttle could either be fully or partially attributed [1]. Therefore, the capability of the accurate prediction of the thermal loading and heat transfer rates is crucial to reduce the uncertainty in the design of rocket engines. In this work, we develop a flamelet-based combustion model for the prediction of convective wall-heat transfer. This model is an extension of the flamelet/progress variable model (FPV) [2], in which an adiabatic combustion process is assumed. The FPV model is capable of representing the detailed flame-structure in terms of a reduced set of describing variables (namely mixture fraction and a progress variable). The turbulence/chemistry interaction is modeled by a presumed probability density function (PDF). The resulting reduction in computational complexity makes this formulation attractive for practical applications. To incorporate non-adiabatic process into the steady-state flamelet formulation, extensions have been proposed [3, 4, 5, 6, 7] for both radiative and convective heat transfer. In these models, heat-loss is introduced globally either through a point-local source term or a prescribed enthalpydefect profile, which increases the dimensionality of the manifold. Therefore, an additional reaction coordinate is introduced (enthalpy or temperature). 1

In this work, the convective heat-loss is introduced into the flamelet formulation locally (which is consistent with the physical space), and its impact on the flame is naturally captured through the flamelet equation. Convective heat-loss was estimated by Lee et al. [7] to be approximately three times as important as the radiative heat-loss under a typical operating condition for rocket engine applications. Therefore, effects of radiative heat-loss are not considered in this work. The mathematical model is described in the next section. This non-adiabatic model is then applied in a direct numerical simulation (DNS) simulation and results are discussed in Sec. 3. The paper finishes with conclusions. 2. Mathematical Model 2.1 Governing Equations The instantaneous conservation equations for mass, momentum, mixture fraction, species mass fraction, and temperature can be written in dimensionless form as: D t = r u, D t u = rp + 1 Re r, D t Z = 1 Re Sc r ( rz), D t Y = 1 Re Sc r ( ry)+ṁ, c p D t T = 1 Re Sc r ( rt )+ Q H 1 Re Sc + Ec Re : ru +EcD tp, (1e) in which u is the velocity vector, is the viscous stress tensor, and D t = @ t + u r is the substantial derivative. The mixture fraction is denoted by Z, species mass fraction is denoted by Y, and temperature is denoted by T. In addition, α is the species diffusivity (which is here assumed to be equal for all species), c! is the specific heat at a constant pressure, λ is the heat conductivity, v! is the diffusion velocity of species i, and Q! is the heat release rate. 2.2 Flamelet Formulation In the laminar flamelet model, a turbulent diffusion flame is modeled as an ensemble of laminar flamelets [8, 9]. This consideration corresponds to the asymptotic behavior of the flame at sufficiently large Damköhler number or sufficiently high activation energy, where chemical reactions and heat transfer occur in a thin layer. In addition, if the characteristic length scale of the flame is smaller than that of the surrounding turbulence, the turbulent eddies cannot penetrate the reaction zone. The effect of turbulence in this so-called flamelet regime results in a deformation and stretching of the flame sheet, such that the flame itself can still be considered as laminar. Under the aforementioned condition, the direction normal to the flame surface is associated with the mixture fraction, while spatial changes along the other directions are negligible. Thus, the one-dimensional laminar flamelet equations can be derived in mixture fraction space. For unity- Lewis number, they can be written as: X i c p,i Y i v i! rt (1a) (1b) (1c) (1d) 2

@ t Y Z 2 @ 2 T @Z 2 Z 2 @ 2 Y @Z 2 = ṁ, 1 @c p c p @Z @T @Z = Q H c p, @ t T Z 2 (2b) and the scalar dissipation rate is defined as Z =2 rz rz. (3) For the steady-state flamelet model, the temporal derivatives in Eqs. (2) are neglected under the assumptions that the chemical time scales are short compared to those associated with the hydrodynamic time scale. For simulations of reactive turbulent flows, the solutions to Eqs. (2) are pre-computed with modeled scalar-dissipation rate profiles (Z) = St exp 2 [erfc 1 (2Z st )] 2 [erfc 1 (2Z)] 2, (4) with respect to different values of χ!". Dirichlet boundary conditions are set at Z = 1 (fuel stream) and Z = 0 (oxidizer stream) for both temperature and species mass fractions. The flame conducts no heat-or-mass-transfer in addition to those through the boundaries. Therefore, this configuration is referred to as an adiabatic formulation. 2.3 Modeling of Heat-Loss Effects in the Flamelet Formulation In the following, we are concerned with the modeling of wall-heat loss effects and flame/wall interaction in the context of the laminar flamelet formulation. In this framework, the direction of heat flux is assumed to be primarily aligned with the gradient of mixture fraction, i.e. normal to the flame surface. As a result, the effect of heat-loss can be introduced by an additional thermal boundary condition, whose location in the mixture fraction coordinate is demoted by Z!"##. To illustrate this, consider Fig. 1, showing the schematic corresponding to a counter-flow diffusion flame. (2a) Oxidizer x Flamelet Q C T Flame Thermal BC ZWall, xwall Fuel Figure 1: Schematic of counter-flow diffusion flame corresponding flame-normal heat transfer in which the heat-flux vector is aligned with the mixture-fraction gradient. 3

The thermal boundary condition can be either of Dirichlet-type: T (Z = Z wall )=T wall (5) or Neumann-type: Q C = Z 2 ( @T @Z @T Z=Z wall @Z Z=Z + ), wall (6) where T!"## is the wall temperature and Q! is the heat transfer rate at the wall. Without loss of generality, we assume in this study that the wall approaches from the fuel-rich side, i.e. Z!"## > Z!". The source terms in Eqs. (2) are set to zero for Z (Z!"##, 1), if an inert wall is assumed. In summary, the non-adiabatic flamelet model consists of the following set of flamelet equations: @ t Y Z @ 2 Y 2 @Z 2 = ṁ, Z2 (0, Z wall), (7a) @ t T Z @ 2 T 2 @Z 2 Z 1 @c p @T 2 c p @Z @Z = Q H,Z2(0, Z wall ), c p (7b) @ t Y Z @ 2 Y 2 @Z 2 =0,Z2 (Z wall, 1), (7c) @ t T Z 2 @ 2 T @Z 2 Z 2 1 @c p c p @Z @T @Z =0,Z2 (Z wall, 1), subject to the corresponding boundary conditions: Y(Z = 0) = Y oxidizer, Y(Z = 1) = Y fuel, Y(Z = Z wall )=Y(Z = Z + wall ),@ ZY(Z = Z wall )=@ Z Y(Z = Z + wall ), T (Z = 0) = T oxidizer,t(z = 1) = T fuel, T (Z = Z wall )=T wall. The non-adiabatic flamelet model proposed contains three different time scales, which are the characteristic diffusion time scale τ!"##, the convective heat-transfer time scale τ!"#$, and the characteristic chemical time scale τ!"#$. The convective time scale is not explicitly specified but embedded in the time scale for the change of Z!"##. The ratio between them scales as [7, 10] di : conv : chem =Da: DaPr Nu :1, (9) where Da is the second Damköhler number, Pr is the Prandtl number, and Nu is the Nusselt number. For the combustion process in rocket engines, the first two terms are both much greater than one. The time-scale separation ensures that the flame-structure reaches the steady state under the consideration of transient effects of the convective heat-loss and limits the influence of the convective time scale to the heat-loss rate only. It is thus justified that Eqs. (7) can be solved under the steady-state condition. (7d) (8a) (8b) (8c) (8d) For a user-specified wall temperature T!"##, the steady-state solutions to Eqs. (7) are obtained with respect to a variation in of χ!" and Z!"##, which generates a three-dimensional manifold. This manifold is then parameterized by the mixture fraction, progress variable, and temperature. This expression can be written in the following form: 4

= (Z, C, T). (10) Correspondingly, the transport equations of these three quantities are solved in the CFD code for simulations utilizing this model. Thus, all the thermo-chemical information can be retrieved from the pre-computed table defined in Eq. (10). A series of sample flamelet solutions with respect to various wall locations Z!"## at a fixed scalar dissipation rate of χ!" = 0.1 (s!! ) are shown in Fig. 2. The operating condition is listed in Table (1). The figure on the left shows that the temperature decreases monotonically as the wall is approaching to flame. The sudden change of the slope at Z!"## indicates a local heat-loss to the wall. The figure on the right shows the corresponding profile of Y!!!, which is less influence by the convective heat-loss to the wall. 2200 1 Temperature, T [K] 2000 1800 1600 1400 1200 1000 800 600 400 Mass Fraction of H 2 O, Y H2O 0.9 0.8 0.7 0.6 0.5 200 0 0.2 0.4 0.6 0.8 1 Mixture Fraction, Z (a) 0.4 0 0.2 0.4 0.6 0.8 1 Mixture Fraction, Z (b) Figure 2: Non-adiabatic flamelet solutions of (a) temperature and (b) mass fraction of H! O for various wall locations Z!"## and a constant scalar dissipation rate ofχ!" = 0.1 s!!. Table 1: Operating conditions and thermochemical parameters of the DNS for non-premixed flame-wall interaction. Fuel-Stream Oxidizer-Stream Mixture Y!! = 0.2, Y!!! = 0.2 Y!! = 0.6, Y!!! = 0.4 Temperature [K] 300 300 Wall Temperature [K] 300 Stoichiometric Mixture 0.2743 Pressure [bar] 1 Prefactor, A [mol, cm, sec] 2.25 10!" Activation Energy, E! [J/mol] 71,128! Heat of combustion, ΔH! [J] 483.67 10! Heat Capacity, c! [J/(gK)] 3.5 5

3. A Priori Study of the Non-adiabatic FPV-model Using DNS 3.1 Problem Formulation The present DNS is conducted as an a priori study of the proposed non-adiabatic FPV-model. The objective is to examine the effect of the conductive wall heat flux on the flame structure and to evaluate the performance of the model in a configuration corresponding to a H 2 /O 2 diffusion flame stabilized in the near-wall region. While the non-adiabatic FPV-model is equipped with the capability of a detailed chemical mechanism as well as the radiative heat transfer, the present DNS study uses a single-step reaction model and neglects the radiation to focus on the convective heat-loss to the wall. The single-step irreversible reaction is described as: 2H 2 +O 2! 2H 2 O. (11) The corresponding species and heat source terms in Eq. (1d) and Eq. (1e) are ṁ =(W w) D ref U ref ref, (12a) Q H =( HC 0 w) D ref U ref ref c p T ref, (12b) and w =( Y H 2 W H2 ) 2 ( Y O 2 W O2 )Ae E a RT [ 1 cm 3 s ]. (13) Other choices on the modeling parameters and assumptions include a constant specific heat, a temperature-dependent dynamic viscosity, a constant Prandtl number of Pr = 0.708, an idealgas equation of state, and unity Lewis numbers. The impact of these assumptions and choices are discussed by Wang and Trouvé [11], where a similar model and configuration were chosen for their DNS of non-premixed flame-wall interactions. All thermochemical parameters are summarized in Table 1. Figure 3: Schematic of the DNS configuration for a turbulent H 2 /O 2 diffusion flame stabilized near a solid wall. 6

The schematic of the DNS configuration is shown in Fig. 3. The simulation is conducted in a three-dimensional, spanwise-periodic box, with an inflow boundary condition at x/d!"# = 0, a wall boundary at y/d!"# = 0, and nonreflecting boundaries at x/d!"# = 20 and y/d!"# = 8. The wall boundary for the temperature is either Dirichlet (non-adiabatic cases) or Neumann (adiabatic cases). The non-premixed combustion is over-ventilated with the fuel stream injected at y-locations below D!"# = 0.5 cm and oxidizer stream injected above it. The inflow profile is a synthetic turbulent flow, of which the mean flow obeys the law of the wall [22]. The turbulent perturbation is generated using the method of [23]. The computational domain consists of 8.2 million grid points. The y-direction is discretized with 320 grid points following an exponential growth rate, while x- and y- directions are uniformly discretized by 400 and 64 grid points respectively. 3.2 Results Table 2: Set-up of the DNS and non-adiabatic FPV cases Case Notation Comments DNS-Adiab DNS, adiabatic at the solid wall DNS-WHT DNS, isothermal at the solid wall FPV-Adiab Flamelet/progress variable model, adiabatic formulation FPV-WHT Flamelet/progress variable model, wall-heat loss formulation In this section, DNS results will be presented and compared. To this end, four computations are performed. Two DNS simulation (DNS-Adiab and DNS-WHT) with different thermal boundary conditions are performed to access the effects of wall-heat transfer on the flame. In these two calculations, the transport equations of species mass fractions and temperature are solved. Then, two flamelet-model calculations (FPV-Adiab and FPV-WHT) with different formulations are conducted to assess of the non-adiabatic flamelet model developed in this work. The flamelet calculations only solve the transport equations of the corresponding describing variables. All of the four cases are well resolved and no turbulence closure model is used. For reference, the case notations are listed in Tab. 2. (a) (b) Figure 4: Comparison of mean temperature field obtained from (a) DNS-Adiab and (b) DNS- WHT. For DNS-WHT, the wall temperature is set to be 300 K. The solid black lines represent the stoichiometric condition. 7

Comparison of mean temperature fields between DNS-Adiab and DNS-WHT are shown in Fig. 4. It can be seen that the change of the thermal boundary condition at the wall introduces significant heat-loss. The flame quenches, for the non-adiabatic case, in the thermo boundary layer. The mean temperature along the stoichiometric line is also reduced. The dilatation is attenuated due to the lower flame temperature in the non-adiabatic case, thereby shifting the stoichiometric line closer to the wall. The considerable effects of the wall-heat loss on the flame/flow-field structure can be accurately described by the non-adiabatic flamelet formulation. Figure 5 illustrates comparison of mean temperature fields between FPV-Adiab and FPV-WHT, where FPV-Adiab utilizes the original adiabatic formulation of the FPV model and FPV-WHT adapts the non-adiabatic formulation introduce in this work. The non-adiabatic FPV model captures the effect of the heat-loss on the flame (including the quenching) and produces an almost identical mean temperature profile. (a) (b) Figure 5: Comparison of mean temperature field obtained from (a) FPV-Adiab and (b) FPV- WHT. For FPV-WHT, the wall temperature is set to T!"## = 300 K. The solid black lines represent the stoichiometric condition. (a) (b) Figure 6: Comparison of (a) mean temperature and (b) mass fraction of H! O obtained at different streamwise location between DNS-WHT, FPV-WHT, and FPV-Adiab. 8

More quantitative comparisons are presented in Fig. (6), showing wall-normal profiles of the mean temperature and the mass fraction of H! O at different streamwise locations are illustrated. The results are obtained from DNS-WHT, FPV-WHT, and FPV-Adiab. For both quantities, the non-adiabatic FPV model gives very accurate prediction compared with the DNS results. The effect of heat-loss is observed to be more significant for temperature than Y!!!. The impact becomes stronger with increasing downstream position and the amount of convective heat-loss accumulates, indicated by the larger deviation between the adiabatic and non-adiabatic FPV models. We also compare the heat flux at the wall predicted by FPV-WHT and DNS-WHT. The accurate prediction of the heat-flux is essential for engineering applications. As shown in Fig. 7, the maximum heat flux is about 110 kw/m 2 and appears at x/l=8. The non-adiabatic FPV formulation accurately predicts the entire heat flux profile, and deviation with the DNS-results are less that 5 pervent. Figure 7: Comparison of heat flux at the wall obtained from DNS-WHT and FPV-WHT. 4. Conclusions A turbulent combustion model for predicting wall-heat transfer and describing flame/wall interaction has been developed. This model extends the adiabatic flamelet/progress variable model by introducing an addition thermal boundary condition. Using this formulation, the solution obtained from the steady non-adiabatic flamelet equations are parameterized in terms of mixture fraction, progress variable, and temperature. Compare to previously developed models, the present non-adiabatic formulation is selfcontained within the flamelet framework, thus offering the advantage that the model does not require additional parameters or user-specified enthalpy profile. Flame quenching is also naturally incorporated in this formulation. 9

The non-adiabatic model was applied to a resolved simulation of a turbulent H2/O2 diffusion flame that is stabilized at an inert isothermal wall. The model prediction is then compared with DNS results. The comparison suggests that the present non-adiabatic formulation provides an adequate description of convective heat-loss effect. An accurate prediction is obtained for the mean scalar fields of temperature and major species, as well as for wall-heat flux profile, which is of major engineering interest. The present formulation can also be extended to LES calculations via a presumed PDF closure. Application of this approach to confined combustors is the topic of ongoing research. 5. Acknowledgements Financial support through NASA and Stanford School of Engineering Fellowship is gratefully acknowledged. 6. References [1] J. C. Monk and O. K. Goetz. Combustion device failures during space shuttle main engine development, 5th International Symposium on Liquid Space Propulsion Long Life Combustion Devices Technology, Chattanooga, TN, 2003 [2] C. D. Pierce and P. Moin, J. Fluid Mech. 504 (2004) 73-97. [3] B. Marracino and D. Lentini, Combust. Sci. Tech. 128.1-6 (1997) 23-48. [4] M. Hossain, J. C. Jones, W. Malalasekera, Flow, Turb. Combust. 67.3 (2001) 217-234. [5] S. Navarro-Martinez, A. Kronenburg, F. Di Mare, Flow, Turb. Combust. 75.1-4 (2005) 245-274. [6] L. Shunn, Large-Eddy Simulation of Combustion Systems with Convective Heat-Loss. PhD thesis, Stanford University, 2009. [7] D. J. Lee, S. Thakur, J. Wright, M. Ihme, W. Shyy, AIAA Paper 6125 (2011). [8] N. Peters, Prog. Energy Combust. Sci. 10.3 (1984) 319-339. [9] N. Peters, Combust. Sci. Tech. 30.1-6 (1983) 1-17. [10] L. Vervisch, T. Poinsot, Annual Rev. Fluid Mech. 30.1 (1998) 655-691. [11] Y. Wang, A. Trouvé, Combus. Flame 144.3 (2006) 461-475. 10