HOT SURFACE IGNITION OF HYDROGEN-AIR MIXTURES

Similar documents
Hot Surface Ignition and Flow Separation

Hot surface ignition of stoichiometric hydrogen-air mixtures

Effect of orientation on the ignition of stoichiometric ethylene mixtures by stationary hot surfaces

Ignition of hydrogen-air mixtures by a localized stationary hot surface

Hot Surface Ignition of Ethylene-Air mixtures: Selection of Reaction Models for CFD Simulations

Laminar Premixed Flames: Flame Structure

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

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

CFD SIMULATION OF HYDROGEN RELEASE, DISPERSION AND AUTO-IGNITION IN ENCLOSURES

Ignition of n-hexane-air Mixtures by Moving Hot Spheres

Cyclic Flame Propagation in Premixed Combustion

Ignition Delay Time of Small Hydrocarbons-Nitrous Oxide(-Oxygen) Mixtures

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

Lecture 7 Flame Extinction and Flamability Limits

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

On two fractional step finite volume and finite element schemes for reactive low Mach number flows

Ignition. Jerry Seitzman. Temperature (K) School of Aerospace Engineering Review. stable/steady self-sustained propagation of premixed flames

Cellular structure of detonation wave in hydrogen-methane-air mixtures

Shock Wave Boundary Layer Interaction from Reflecting Detonations

DNS of Reacting H 2 /Air Laminar Vortex Rings

Reaction Rate Closure for Turbulent Detonation Propagation through CLEM-LES

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

Dynamics of Excited Hydroxyl Radicals in Hydrogen Based Mixtures Behind Reflected Shock Waves. Supplemental material

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

Asymptotic Structure of Rich Methane-Air Flames

The role of diffusion at shear layers in irregular detonations

Lecture 8 Laminar Diffusion Flames: Diffusion Flamelet Theory

Experimental Study of Minimum Ignition Energy of Lean H 2 -N 2 O Mixtures. Abstract

arxiv: v1 [physics.chem-ph] 6 Oct 2011

Fluid Dynamics and Balance Equations for Reacting Flows

TOPICAL PROBLEMS OF FLUID MECHANICS 97

Presentation Start. Zero Carbon Energy Solutions 4/06/06 10/3/2013:; 1

Modeling and Simulation of Plasma-Assisted Ignition and Combustion

AME 513. " Lecture 8 Premixed flames I: Propagation rates

Combustion. Indian Institute of Science Bangalore

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

Detonation Cell Width Measurements for H 2 N 2 O N 2 O 2 CH 4 NH 3 Mixtures

On the Effect of Variable Opening Geometries, and Operating Conditions on High Pressure Hydrogen Releases: Ignition Risks

APPENDIX A: LAMINAR AND TURBULENT FLAME PROPAGATION IN HYDROGEN AIR STEAM MIXTURES*

EVALUATION OF HYDROGEN, PROPANE AND METHANE-AIR DETONATIONS INSTABILITY AND DETONABILITY

CFD Analysis of Vented Lean Hydrogen Deflagrations in an ISO Container

Hydrogen addition to the Andrussow process for HCN synthesis

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

Formation and Evolution of Distorted Tulip Flames

NUMERICAL SIMULATION OF HYDROGEN COMBUSTION. Jan-patrice SIMONEAU, FRAMATOME - FRANCE

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

UQ in Reacting Flows

Numerical Simulation of Premixed V-Flame

The Effect of Mixture Fraction on Edge Flame Propagation Speed

A NUMERICAL ANALYSIS OF COMBUSTION PROCESS IN AN AXISYMMETRIC COMBUSTION CHAMBER

Influence of a Dispersed Ignition in the Explosion of Two-Phase Mixtures

Erratum to: High speed mixture fraction and temperature imaging of pulsed, turbulent fuel jets auto igniting in high temperature, vitiated co flows

Numerical simulation study of turbulent combustion phenomena -INTEGRATE Advanced Study Group (ASG)

Numerical investigation of flame propagation in small thermally-participating, adiabatic tubes

Premixed, Nonpremixed and Partially Premixed Flames

Computational Analysis of an Imploding Gas:

10. Buoyancy-driven flow

Lecture 9 Laminar Diffusion Flame Configurations

REDIM reduced modeling of quenching at a cold inert wall with detailed transport and different mechanisms

DYNAMIC LOAD ANALYSIS OF EXPLOSION IN INHOMOGENEOUS HYDROGEN-AIR

Combustion Chemistry

THERMODYNAMIC ANALYSIS OF COMBUSTION PROCESSES FOR PROPULSION SYSTEMS

Measurement and chemical kinetic model predictions of detonation cell size in methanol-oxygen mixtures

Mauro Valorani Dipartimento di Meccanica e Aeronautica, University of Rome

Experimental study of the combustion properties of methane/hydrogen mixtures Gersen, Sander

Workshop Paris, November 2006

CFD study of gas mixing efficiency and comparisons with experimental data

Chapter 1. Governing Equations of GFD. 1.1 Mass continuity

Combustion Theory and Applications in CFD

CHARACTERISTIC OF VORTEX IN A MIXING LAYER FORMED AT NOZZLE PITZDAILY USING OPENFOAM

Ignition delay-time behind reflected shock waves of small hydrocarbons nitrous oxide( oxygen) mixtures

Initiation of stabilized detonations by projectiles

COMBUSTION CHEMISTRY COMBUSTION AND FUELS

Opposed Flow Impact on Flame Spread Above Liquid Fuel Pools

arxiv: v1 [physics.flu-dyn] 25 Nov 2018

Computational and Experimental Studies of Fluid flow and Heat Transfer in a Calandria Based Reactor

Advanced Numerical Methods for non-premixed Flames

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

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

Statistical Analysis of Electrostatic Spark Ignition

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

DETAILED MODELLING OF SHORT-CONTACT-TIME REACTORS

Detailed and Simplified Chemical Reaction Mechanisms for Detonation Simulation

Application of COMSOL Multiphysics in Transport Phenomena Educational Processes

CHAPTER - 12 THERMODYNAMICS

Rocket Propulsion. Reacting Flow Issues

Application of a Laser Induced Fluorescence Model to the Numerical Simulation of Detonation Waves in Hydrogen-Oxygen-Diluent Mixtures

An artificial nonlinear diffusivity method for shock-capturing in supersonic reacting flows

Mild Ignition Phenomena in Rapid Compression Machines

Thermoacoustic Instabilities Research

Basic concepts in viscous flow

NUMERICAL INVESTIGATION OF IGNITION DELAY TIMES IN A PSR OF GASOLINE FUEL

Introduction to Aerodynamics. Dr. Guven Aerospace Engineer (P.hD)

TAU Extensions for High Enthalpy Flows. Sebastian Karl AS-RF

Detonation Characteristics Of Dimethyl Ether, Methanol and Ethanol Air Mixtures

Jennifer Wen 1 & Changjian Wang 2,3. Warwick Fire, School of Engineering, University of Warwick

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

Propagation of hydrogen-air detonation in tube with obstacles

Well Stirred Reactor Stabilization of flames

Transcription:

HOT SURFACE IGNITION OF HYDROGEN-AIR MIXTURES Melguizo-Gavilanes, J., Mével, R., and Shepherd, J.E. Explosion Dynamics Laboratory Graduate Aerospace Laboratories of the California Institute of Technology (GALCIT) Pasadena, CA, USA, josuemg@caltech.edu ABSTRACT Hot surface ignition is relevant in the context of industrial safety. In the present work, two-dimensional simulations with detailed chemistry, and study of the reaction pathways of the buoyancy-driven flow and ignition of a stoichiometric hydrogen-air mixture by a rapidly heated surface (glowplug) are reported. Experimentally, in hydrocarbon-air, ignition is observed to occur regularly at the top of the glowplug; numerical results for hydrogen-air reproduce this trend, and shed light on this behavior. Flow separation plays a crucial role in creating zones where convective losses are minimized and heat diffusion is maximized, resulting in the critical conditions for ignition to take place. 1. INTRODUCTION Ignition of combustible atmospheres by hot surfaces is a common issue in industrial safety. Determining critical conditions for ignition in terms of surface size and temperature are essential in order to evaluate the potential of an ignition hazard. Classical experimental work on hot surface ignition includes that of Coward and Guest [1], and Kutcha [2]. The former investigated the effect of material (e.g catalytic and non-catalytic surfaces) on ignition thresholds, whereas the latter extended this work to study the effect of variations in size and geometry. The impact of their results was hindered by their inability to measure flow velocity and composition during the ignition event. An extensive review of more recent studies is given by Brabauskas [3]. Experimental work done by Boettcher [4] using a glow plug found the ignition temperature for n-hexane to be essentially insensitive to composition away from flammability limits. Analytical studies have been performed by Gray [5] who investigated the effect of surface to volume ratio, and more recently Laurendeau [6] in which a simple model is proposed to estimate the minimum ignition temperature. Some numerical efforts in this area are due to Kumar [7] who developed a onedimensional model to study hydrogen ignition, and the two-dimensional steady simulations of Adler [8] in which the problem of a circular hot spot in contact with reactive mixture was analyzed. Boettcher [4] carried out simulations to examine predicting lower flammability limits with tabulated chemistry as well as studying the effect of hot surface area on ignition temperature using one-step and detailed chemical reactive models [9]. None of the previous work has been concerned with analyzing in detail the flow field in the vicinity of the hot surface. For an accurate numerical prediction of this flow, it is necessary to solve the conservation equations together with transport of chemical species on a mesh small enough to capture the thermal and hydrodynamic boundary layer surrounding the hot surface. The wide range of temporal and spatial scales involved, as well as the size of detailed chemical kinetic mechanisms pose significant computational

challenges. Hydrogen is one of the fuels for which the chemistry is better known, the detailed kinetic mechanism is of a reasonable size to simulate realistic geometries, and constitutes a good model fuel for understanding the kinetics of complex hydrocarbons. A two-dimensional numerical simulation of the transient viscous flow and ignition of combustible atmospheres using a detailed hydrogen oxidation mechanism is presented. Special attention is given to the near-wall buoyancy flow induced, and flow separation to gain insight on the dynamics, time and location of the ignition event. 2. PHYSICAL MODEL, NUMERICAL APPROACH AND SIMULATION PARAMETERS The heat transfer and ignition process in the gas is modeled by the variable-density reactive Navier- Stokes equations with temperature dependent transport properties. ρ t + (ρu) = (1) (ρu) + (ρuu) = p + τ + ρg t (2) (ρh) + (ρuh) = (κ/c p h) + q chem + Dp t Dt (3) (ρy i ) + (ρuy i ) = (ρd i Y i ) + Ω i t (4) with p = ρ RT, τ = (p + 2 3 µ u)i + µ[ u + ( u)t ] (5) The Sutherland Law, the Eucken Relation and the JANAF polynomials are used to account for the functional temperature dependence of mixture viscosity (µ), thermal conductivity (κ) and specific heat (c p ) respectively. The chemistry is modeled using Mevel s detailed mechanism for hydrogen oxidation which includes 9 species and 21 reactions [1, 11]. In equations (1)-(5), ρ is density, u is the velocity vector, p is pressure, T is the gas temperature, h is the mixture enthalpy, g is the gravitational acceleration, q chem = h i Ω i is the chemical energy conversion rate, Y i is the mass fraction of species, Ω i = ρdy i /dt represents the rate of production/consumption of species, and R is the specific gas constant. The Lewis number is assumed to be unity which results in κ/c p = ρd i, hence, the dynamic thermal diffusivity of species is used to model its mass diffusivity. The equations above are integrated in two dimensions using the Open source Field Operation And Manipulation (OpenFOAM) toolbox [12]. The spatial discretization of the solution domain is done using finite volumes, and the pressure-velocity coupling is achieved using the PIMPLE (PISO+SIMPLE) algorithm. The geometry simulated corresponds to that used in [4], a combustion vessel of 11.4cm x 17.1cm with a glowplug of 9.3mm x 5.1mm located in the center. There are approximately 2k cells in the computational domain, compressed near the wall of the glow plug, with a minimum cell size of 8 µm to allow for enough resolution to resolve the thermal and hydrodynamic boundary layer. Initial conditions are p o = 11 kpa, T o = 3 K, U o = m/s, premixed stoichiometric hydrogen-air mixture (Y H2 =.283, Y O2 =.2264, Y N2 =.7453), with non-slip and adiabatic boundary condition on walls, and a prescribed temperature ramp (spatially uniform) on the surface of the glowplug given by T(t) = T o + rt with r = 22 K/s as in [4]. 3. RESULTS AND DISCUSSION 3.1. Overview, flow structure and temporal evolution In the present work, numerical simulation of ignition of stoichiometric hydrogen-air mixtures by a rapidly heated commerical glow plug previously studied in experiments by Boettcher [4] is performed. 2

In particular, one of the observations made by Boettcher suggested an interesting line of further study. According to his experiments with hydrocarbons (n-hexane), in 53 % of them ignition was observed to occur at the top of the glowplug, 3 % were reported to have taken place on the side of the hot surface, and 17 % in the thermal plume that develops as a consequence of the buoyancy driven flow. Why is ignition more likely to occur at the top? This study aims to answer this question and to understand the physics driving this behavior. t = 2.86 s Velocity (m/s) Velocity (m/s).25.2.15.1.5 -.5.2.15.1.5 -.5 Ux Uy Umag Temperature 1 2 3 4 5 6 1 2 3 4 5 6 Horizontal distance from surface of glowplug (mm) @ y = 5mm 75 6 45 3 75 6 45 3 Figure 1: Left: temperature and velocity (magnitude) field in the vicinity of the glowplug, temperature isocontours and velocity vectors. Right: spatial distribution of velocity and temperature at top y = 9.3mm and side y = 5mm of glowplug. Figure 1 (Left) shows the temperature and velocity (magnitude) fields obtained after 2.86 s of heating, together with temperature isocontours taken every 5 K from 4 K T 93 K, and velocity vectors showing clearly the buoyancy driven flow induced by the glowplug. Figure 1 (Right) shows plots of the spatial distribution of each velocity component, magnitude and temperature at two locations on the glowplug. At the top (y = 9.3 mm) the vertical spatial distribution, and at the side (y = 5 mm) the horizontal spatial distribution. In the vicinity of the hot surface there is development of a thermal boundary layer and thermal plume. Note that in the separated region (top of the glowplug) there is preferential heat diffusion taking place. The thermal plume is delineated by the outermost temperature contour (T = 4 K). The velocity (magnitude) field and velocity vectors show the flow structure near the glowplug. Parcels of fresh cold gas enter the thermal boundary layer from below and heat up slowly as they travel upward in close proximity to the wall. Once they reach the upper right/left corner of the glowplug, the flow separates creating a region at the top of the glow plug where the gas is practically at rest. The gas continues to rise to the top of the combustion vessel, is forced to turn and creates a rather complex vortical flow field, a glimpse of which can be seen on the upper region of the velocity fields. Details of this flow field and an unusual cyclic flame oscillation were examined in experiments and simulations by Boettcher et al. [13] for hexane-air mixtures in this geometry. In Fig. 1 (right), the horizontal (U x ) and vertical (U y ) components of the velocity vector, black and red solid lines respectively, magnitude (dashed line) and temperature (blue dashed-dotted line) confirm that at up to.5 mm away from the top surface of the glowplug the flow is essentially stagnant. As a result, the gas can be readily heated up by the hot surface because convective losses are minimal in this region. This plot also shows the temperature distribution of the thermal plume up to 6 mm away from the glowplug surface. At the side of the glowplug (y = 5 mm), bottom plot on Fig. 1 (Right), the temperature and velocity magnitude plots show the thermal and hydrodynamic boundary layer thickness, 5.5 mm and 4 mm respectively. The negative values of U x (gas moving left) display how parcels of fluid 3

are brought into the thermal boundary layer from colder regions away from the glowplug, slowed as they approach the hot surface, changing direction gradually (see increase in U y ), subsequently reaching a maximum, immediately followed by a decrease to zero velocity at the wall consistent with the non-slip condition imposed. 27 24 21 18 15 12 Glowplug surface temperature T max in computational domain 27 24 21 18 15 12 2.864 2.8645 2.865 6 3.25.5.75 1 1.25 1.5 1.75 2 2.25 2.5 2.75 3 time (s) Figure 2: Temperature maximum in computational domain during the course of the simulation. Inset: closeup to ignition event. To accurately determine the ignition time, τ ign, the temperature maximum in the computational domain and glowplug surface temperature are monitored during the simulation. The inset in Fig. 2 shows a close up of the main heat release event. The time to ignition is τ ign = 2.864435 s. The temperature of the glowplug surface is 93 K at this time. The strong dependence of ignition threshold on temperature is consistent with constant volume explosion simulations for stoichiometric hydrogen-air mixtures. These simulations show a rapid increase in ignition delay times below 1 K, indicative of a sharp change in activation energy in this temperature range. For instance, the constant volume ignition delay time at K is 94 ms decreasing to 2.9 ms at 93 K. 3.2. Ignition evolution Figure 3 shows velocity, temperature and product (H 2 O) mass fraction fields together with velocity vectors at four instances during the simulation. At t = 2.86432 s chemical activity is already taking place at the top of the glowplug where the temperature is highest, and convective losses are minimal. The temperature maximum in the domain (T = 93 K) corresponds to that of the glowplug surface until ignition takes place. At t = 2.864435 s, 115 µs later, an ignition center appears on the temperature field as closed contours at the top surface of the glowplug. The temperature contours are rescaled to cover the full range of temperature within the computational domain at each time shown in Fig. 3. This ignition center is accompanied by an associated localized increase in H 2 O concentration and gas velocity. The velocity vectors show how the ignition kernel abruptly pushes away the surrounding gas. Further acceleration of the gas from 2 m/s to 1 m/s in 28 µs can be seen on the velocity fields at t = 2.864463 s. The rapid expansion is evidenced by the size of the velocity vectors. A nascent flame is observed in the temperature contours, and the fuel is nearly completely consumed within the flame kernel, as the mass fraction of H 2 O is.21 close to the theoretical value of.25 given by complete oxidation of a stoichiometric mixture. The last frame shows the early stages of flame propagation, the gas continues to be pushed radially outwards very rapidly, and the shape of the flame is determined by the preferential propagation of the combustion front along the thermal plume where the fresh combustible mixture is 4

t = 2.86432 s t = 2.864435 s t = 2.864463 s t = 2.864577 s Figure 3: Velocity, product mass fraction and temperature fields. Top Left: at t = 2.86432 s - shortly before ignition. Top Right: at t = 2.864435 s - ignition event. Bottom Left: at t = 2.864463 s - shortly after ignition/flame kernel formation. Bottom right: at t = 2.864577 s - early stages of flame propagation. hottest. Similar behavior was observed and simulated by Boettcher et al. [13] for the ignition of hexaneair mixtures. The temperature contours show clearly, a nearly uniform high temperature region within the flame. 3.3. Energy equation analysis To gain additional insight into the processes taking place at the top of the glowplug, each of the terms in the energy conservation equation is plotted along the vertical centerline from the surface of the glowplug (see Figure 4). The plots are taken at the same times as in Figure 3 to allow for a direct comparison. The abscissas represent the normal vertical distance from the surface of the glowplug, whereas the ordinates show the corresponding energy density and temperature. The solid lines are the convective and diffusive heat losses, and the chemical source term given respectively by hconvection = (ρuh), hdi f f usion = (κ/c p h), and hs ource = qchem. The dashed line is the sum of the above terms, and the dashed-dotted line is the temperature. Shortly before ignition (Fig. 4 top left), close to the glowplug surface, the source term is mostly balanced by diffusion. The dip in the convective term is due to the expansion of the gas taking place in this area as a result of the initial heat release produced by the chemistry; the sum is positive up to.5 mm from the glowplug surface, and the temperature maximum 5

hterms x 1-9 (W/m 3 ) @ t = 2.86432s.3.15 -.15 1 8 7 -.3 6 1 2 3 4 5 hterms x 1-9 (W/m 3 ) @ t = 2.864435s 26 13-13 Temperature hconvection hdiffusion hsource Sum 12 11 1 8 7-26 6 1 2 3 4 5 hterms x 1-9 (W/m 3 ) @ t = 2.864463s 26 13-13 24 21 18 15 12-26 6 1 2 3 4 5 hterms x 1-9 (W/m 3 ) @ t = 2.864577s 26 13-13 27 24 21 18 15 12-26 6 1 2 3 4 5 6 Figure 4: Ignition evolution: contributions of each term in energy equation and temperature along normal distance from top surface of the glowplug. Top Left: at t = 2.86432 s - shortly before ignition. Top Right: at t = 2.864435 s - ignition event. Bottom Left: at t = 2.864463 s - shortly after ignition/flame kernel formation. Bottom right: at t = 2.864577 s - early stages of flame propagation. remains at the wall. Further away from the glowplug s wall however (.5 5 mm), convection balances diffusion. In Fig. 4 top right, 115 µs later, the temperature maximum is no longer at the wall but roughly.12 mm away from the surface of the glowplug, hence, the rate at which heat is diffused back to the wall is not large enough to counteract the rate at which heat is released by the chemistry at this location, signaling the birth of an ignition center. The explosive nature of the ignition event can be visualized in the increase of over 1 times in the source term over 115 µs. The bottom left plot of Fig. 4 shows the structure of the ignition center and the birth of an expanding flame kernel. Due to the abrupt expansion of the gas, and associated velocity (1 m/s), the chemical source term is balanced at the flame front (.5 mm) mostly by the convective term. At the wall, the balance is maintained by diffusion as expected. The plot at the bottom right of Fig. 4, displays very clearly the structure of the flame propagating away from the surface of the glowplug. 3.4. Chemical pathways Plotting the contributions of each term in the energy equation and temperature perpendicularly from the top surface of the glowplug was necessary to find the exact location where ignition takes place, y =.12 mm on the centerline above the glowplug. To investigate the chemistry in more detail, temporal probes were collected for temperature, chemical source term, diffusive and convective losses, together with species profiles. The plots in Figure 5, show the start of heat release as early as t = 2.86435 s, diffusion 6

and convection immediately counteract the source term. The ignition event is marked by the sudden increase in the rate at which energy is deposited in the gas, rise in temperature, rapid consumption of fuel and the production of H 2 O and reactive transient species, H, OH and O. Before ignition takes place, at t = 2.863 s the species H 2 O 2 and HO 2 start to build up. At t = 2.864 s, HO 2 increases substantially just before ignition occurs. Figure 6 displays the reaction pathway diagram. It is seen that the chain branching reactions, R 1 : H + O 2 = OH + O and R 2 : O + H 2 = OH + H, account only for 15 to 2 % of the reactants consumption. Molecular hydrogen and oxygen are consumed by R 3 : H 2 + OH = H 2 O + H and R 4 : H + O 2 (+M) = HO 2 (+M), respectively. While R 3 is highly exothermic and induces significant release of energy, R 4 delays the formation of OH radical. Only 34 % of OH is formed by R 1 and R 2, whereas the sequence R 4 followed by R 5 : HO 2 + H = OH + OH represents 66 % of OH formation. Under this low temperature conditions, indirect formation of OH radicals through linear chain chemical processes dominates over direct, chain branching reactions. This is consistent with the expected and previously observed behavior of low-temperature homogeneous reaction kinetics. hterms x 1-9 (W/m 3 ) 26 13-13 -26 2.8643 2.8644 2.8645 time (s) 18 17 16 15 14 13 12 11 1 Y i - Major Species.3.25.2.15.1.5 2.8643 2.8644 2.8645 time (s) 18 17 16 15 14 13 12 11 1 Y i - Minor Species 2 1.5 1.5 18 17 16 15 14 13 12 11 1 2.863 2.8635 2.864 2.8645 time (s) hconvection hdiffusion hsource Sum Temperature H2 x 1 O2 H2O H x 1 O x 1 OH x 1 H2O2 x 5 HO2 x 1 Figure 5: Temporal evolution of each term in energy equation, temperature and species mass fractions at the ignition location, y =.12 mm. 4. CONCLUSION Two-dimensional simulations of ignition by a transiently heated commercial glowplug were performed. In agreement with experiments with hydrocarbon-air mixtures, ignition was observed to occur at the top of the glowplug. The details of the ignition kernel evolution was explained by means of velocity, product mass fraction and temperature fields. Additional insight was achieved by analyzing the individual contributions of the terms in the energy conservation equation. Close to the wall, diffusion counteracts 7

H2 O2 84 16 OH O H (2) M (8) (89) H2O+H H+OH (9) (59) OH+O HO2+M (2) (84) (14) (16) (68) (1) OH H O HO2 (11) (28) (84) H+M (66) H (32) (72) (15) H2O+M OH+OH H2O+O (1) (6) H2O Figure 6: Reaction pathway analysis at the ignition location, y =.12 mm. Boxes represent species reservoirs, solid lines are reservoir inputs, and dashed lines are reservoirs outputs. the heat release due to the chemistry, whereas far away, convection and diffusion maintain the balance. Ignition occurs when the heat release rate is greater than the rate at which heat is diffused back to the wall. Results show the importance of flow separation in creating zones that are prone to ignition. Under these thermodynamic conditions, the reaction pathway analysis showed that ignition is essentially driven by a linear chain chemical process. The chain branching reactions, H + O 2 = OH + O and O + H 2 = OH + H constitute minor pathways in producing reactive radicals. The main sequence leading to hydroxyl radical formation is: H + O 2 (+M) = HO 2 + M; HO 2 + H = OH + OH. We anticipate that analogous and much more complex processes also dominate the thermal ignition chemistry of hydrocarbons in this configuration. ACKNOWLEDGMENTS This work was carried out in the Explosion Dynamics Laboratory of the California Institute of Technology, J. Melguizo-Gavilanes was supported by the Natural Sciences and Engineering Research Council of Canada (NSERC) Postdoctoral Fellowship Program, and R. Mével by The Boeing Company through a Strategic Research and Development Relationship Agreement CT-BA-GTA-1. 8

REFERENCES 1. Coward, H.F. and Guest, P.G., Ignition of natural gas-air mixtures by heated metal bars 1, Journal of the American Chemistry Society, 1927, 49(1), pp. 2479-2486 2. Kuchta, J.M., Investigation of fire and explosion accidents in the chemical, mining, and fuel related industries, 1985, Bulletin 68, Bureau of Mines. 3. Brabrauskas, V., Ignition Handbook, 23, Fire Science Publishers. 4. Boettcher, P.A., Thermal Ignition, 212, Ph.D Thesis. Caltech. 5. Gray, B.F., The dependence of spontaneous ignition temperature on surface to volume ratio in static systems for fuels showing negative temperature coefficient, Combustion and Flame, 197, 14(1), pp. 113-115 6. Laurendeau, N.M., Thermal ignition of methane-air mixtures by hot surfaces: A critical examination, Combustion and Flame, 1982, 46, pp. 29-49 7. Kumar, R.K., (1989). Ignition of hydrogen-oxygen-diluent mixtures adjacent to a hot, nonreactive surface, Combustion and Flame, 1989, 75(2), pp. 197-215 8. Adler, J., Ignition of a combustible stagnant gas layer by a circular hot spot. Combustion Theory and Modeling, 1999, 3(2), pp. 359-369 9. Boettcher, P.A., Mével, R., Thomas, V. and Shepherd, J.E., (212). The effect of heating rates on low temperature hexane combustion, Fuel, 212, 96, pp. 392-43 1. Mével, R., Javoy, S., Lafosse, F., Chaumeix, N., Dupré, G. and Paillard, C.E., Hydrogen-nitrous oxide delay time: shock tube experimental study and kinetic modelling, Proceedings of The Combustion Institute, 29, 32, pp. 359-366. 11. Mével, R., Javoy, S. and Dupré, G., A chemical kinetic study of the oxidation of silane by nitrous oxide, nitric oxide and oxygen, Proceedings of The Combustion Institute, 211, 33, pp. 485-492. 12. Weller, H.G., Tabor, G., Jasak, H., and Fureby, C., A tensorial approach to continuum mechanics using object-oriented techniques, Journal of Computational Physics, 1998, 12, pp. 62-631 13. Boettcher, Menon, S.K., Ventura, B.L., Blanquart, G. and Shepherd, J.E., Cyclic Flame Propagation in Premixed Combustion, Journal of Fluid Mechanics, 213, 735, pp. 176-22 9