Effects of radiative heat loss on the extinction of counterflow premixed H 2 air flames

Similar documents
Yiguang Ju, Hongsheng Guo, Kaoru Maruta and Takashi Niioka. Institute of Fluid Science, Tohoku University, ABSTRACT

Lecture 8 Laminar Diffusion Flames: Diffusion Flamelet Theory

Effects of radiative emission and absorption on the propagation and extinction of premixed gas flames

Lecture 7 Flame Extinction and Flamability Limits

Laminar Premixed Flames: Flame Structure

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

HOT PARTICLE IGNITION OF METHANE FLAMES

Premixed, Nonpremixed and Partially Premixed Flames

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

EFFECTS OF INERT DUST CLOUDS ON THE EXTINCTION OF STRAINED, LAMINAR FLAMES AT NORMAL- AND MICRO-GRAVITY

The Effect of Mixture Fraction on Edge Flame Propagation Speed

Extinction Limits of Premixed Combustion Assisted by Catalytic Reaction in a Stagnation-Point Flow

Numerical evaluation of NO x mechanisms in methane-air counterflow premixed flames

a 16 It involves a change of laminar burning velocity, widening or narrowing combustion limits for

COMPUTATIONAL SIMULATION OF SPHERICAL DIFFUSION FLAME DYNAMICS AND EXTINCTION IN MICROGRAVITY TIANYING XIA. A thesis submitted to the

EFFECTS OF PRESSURE AND PREHEAT ON SUPER-ADIABATIC FLAME TEMPERATURES IN RICH PREMIXED METHANE/AIR FLAMES

Structures of Turbulent Bunsen Flames in the Corrugated-Flamelet Regime

CALCULATION OF THE UPPER EXPLOSION LIMIT OF METHANE-AIR MIXTURES AT ELEVATED PRESSURES AND TEMPERATURES

Asymptotic Structure of Rich Methane-Air Flames

Online publication date: 03 February 2010

Effects of turbulence and flame instability on flame front evolution

Asymptotic Analysis of the Structure of Moderately Rich Methane-Air Flames

Ignition and Extinction Fronts in Counterflowing Premixed Reactive Gases

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

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

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

Period Doubling Cascade in Diffusion Flames

Analysis of homogeneous combustion in Monolithic structures

Modelling of transient stretched laminar flame speed of hydrogen-air mixtures using combustion kinetics

Stability Diagram for Lift-Off and Blowout of a Round Jet Laminar Diffusion Flame

EXPERIMENTAL AND NUMERICAL STUDIES FOR FLAME SPREAD OVER A FINITE-LENGTH PMMA WITH RADIATION EFFECT

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

Premixed Flame Extinction by Inert Particles in Normal- and Micro-Gravity

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

Abstract It is of great theoretical interest to investigate the interaction between flames and vortices as a model problem to study turbulent combusti

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

Nonpremixed-gas flames - counterflow. Research, TwentySeventh International Symposium on Combustion,

FLAME AND EDDY STRUCTURES IN HYDROGEN AIR TURBULENT JET PREMIXED FLAME

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

Combustion. Indian Institute of Science Bangalore

Fluid Dynamics and Balance Equations for Reacting Flows

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

Modeling ion and electron profiles in methane-oxygen counterflow diffusion flames

Predicting NO Formation with Flamelet Generated Manifolds

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

Microgravity Burner-Generated Spherical Diffusion Flames: Experiment and Computation

Super-adiabatic flame temperatures in premixed methane-oxygen flames

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

Chemically-Passive Suppression of Laminar Premixed Hydrogen Flames in Microgravity

Lewis number effects in laminar diffusion flames near and away from extinction

LES Approaches to Combustion

Lower limit of weak flame in a heated channel

Lecture 9 Laminar Diffusion Flame Configurations

Development of Reduced Mechanisms for Numerical Modelling of Turbulent Combustion

TOPICAL PROBLEMS OF FLUID MECHANICS 97

Faculty of Engineering. Contents. Introduction

Effects of Variation of the Flame Area and Natural Damping on Primary Acoustic Instability of Downward Propagating Flames in a Tube

Interaction of Lewis number and heat loss effects for a laminar premixed flame propagating in a channel

Premixed Edge-Flames under Transverse Enthalpy Gradients

Lawrence Berkeley National Laboratory Lawrence Berkeley National Laboratory

Scalar dissipation rate at extinction and the effects of oxygen-enriched combustion

UQ in Reacting Flows

Part II Combustion. Summary. F.A. Williams, T. Takeno, Y. Nakamura and V. Nayagam

Flame Propagation in Poiseuille Flow under Adiabatic Conditions

AME 513. " Lecture 7 Conservation equations

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

Analysis of lift-off height and structure of n-heptane tribrachial flames in laminar jet configuration

Well Stirred Reactor Stabilization of flames

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

EFFECT OF CARBON DIOXIDE, ARGON AND HYDROCARBON FUELS ON THE STABILITY OF HYDROGEN JET FLAMES

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

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

On the critical flame radius and minimum ignition energy for spherical flame initiation

A numerical and experimental investigation of inverse triple flames

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

Dynamics of Premixed Hydrogen-Air Flames in Microchannels with a Wall Temperature Gradient

Thermoacoustic Instabilities Research

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

Direct pore level simulation of premixed gas combustion in porous inert media using detailed chemical kinetics

Numerical modeling of non-adiabatic heat-recirculating combustors

Advanced fundamental topics

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

Turbulent Premixed Combustion

Experimental Study of 2D-Instabilities of Hydrogen Flames in Flat Layers

2D Direct Numerical Simulation of methane/air turbulent premixed flames under high turbulence intensity Julien Savre 04/13/2011

Experimental results on the stabilization of lifted jet diffusion flames

Laminar flame speed (burning velocity) reactants. propagating flame front. products. reactants

SUPPRESSION EFFECTIVENESS OF WATER MIST ON ACCIDENTAL AIRCRAFT FIRES

The role of diffusion at shear layers in irregular detonations

The Seeding of Methane Oxidation

Hydrogen addition to the Andrussow process for HCN synthesis

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

Dynamics and Structure of Dusty Reacting Flows: Inert Particles in Strained, Laminar, Premixed Flames

Overview of Turbulent Reacting Flows

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

Evaluation of Numerical Turbulent Combustion Models Using Flame Speed Measurements from a Recently Developed Fan- Stirred Explosion Vessel

Effect of Varying Composition on Temperature Reconstructions Obtained from Refractive Index Measurements in Flames

Temperature time-history measurements in a shock tube using diode laser absorption of CO 2 near 2.7 µm

Chemical Kinetic Reaction Mechanisms

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

Transcription:

Combust. Theory Modelling 4 (2000) 459 475. Printed in the UK PII: S1364-7830(00)09647-9 Effects of radiative heat loss on the extinction of counterflow premixed H 2 air flames Hongsheng Guo, Yiguang Ju and Takashi Niioka Combustion Research Group, Institute for Chemical Process and Environmental Technology, National Research Council of Canada, Ottawa, Canada K1A 0R6 Department of Engineering Mechanics, Tsinghua University, Beijing 100084, People s Republic of China Institute of Fluid Science, Tohoku University, Katahira, Sendai 980, Japan E-mail: Hongsheng.Guo@nrc.ca Received 16 November 1999 Abstract. Radiation heat loss has an important impact on near-limit flames. It has been shown that radiation heat loss can make stretched CH 4 air flames extinguish at a lower stretch rate. Numerical calculations of counterflow premixed H 2 air flames were conducted using an accurate description of the chemical kinetics and transport properties. Radiation heat loss was considered. The results show that in addition to the stretch extinction limit, radiation heat loss also allows the lower and middle equivalence ratio counterflow premixed H 2 air flames to extinguish at a lower stretch rate. For middle equivalence ratio counterflow H 2 air flames, the closed temperature profiles become distorted O-shaped curves due to the lower Lewis number, being different from those of CH 4 air flames. For higher equivalence ratio counterflow premixed H 2 air flames, there are two stable flame branches a normal flame branch and a weak flame branch. When the equivalence ratio is greater than a critical value, the closed temperature profile curve of every equivalence ratio flame opens and the normal flame curve can be extended to zero stretch rate. The calculation of the concentration limit of one-dimensional planar premixed H 2 air flames was also conducted. The results show that the critical equivalence ratio corresponds to the concentration limit of the one-dimensional planar premixed H 2 air flame. The extension of the flammable region due to the stretch is amplified for the counterflow H 2 air flame because of its much lower Lewis number than that of the counterflow CH 4 air flame. 1. Introduction Flammability limits and the associated limiting mechanisms of premixed flames have been of interest to combustion scientists. For a one-dimensional premixed flame in a doubly infinite domain, it has been well known that radiation heat loss produces a concentration limit [1, 2]. For stretched premixed flames, both experimental and theoretical investigations have shown that an excessive stretch makes the flame extinguish due to incomplete combustion [3, 4]. A proper understanding of the structure and extinction characteristics of stretched laminar flames, such as counterflow flames, is critical to the development and application of the laminar flamelet concept to turbulent flame modelling. Sohrab and Law [5] investigated the effect of radiation heat loss on stretched premixed flames, and concluded that radiation has little impact on the extinction of stretched flames because the stretch rate in their study was not extended to a sufficiently low value. Egolfopoulos Author to whom correspondence should be addressed. 1364-7830/00/040459+17$30.00 2000 IOP Publishing Ltd 459

460 H Guo et al [6] also studied the radiation effects on steady and unsteady stretched flames, and indicated that radiation is important for near-limit stretched laminar flames and for weakly stretched diffusion flames. However, the study was not extended to sufficiently low stretch rates either. A pioneering work by Platt and T ien [7] investigated the extinction of counterflow premixed flames at low stretch rates, and found that the flame extinguishes at a low stretch rate because of the radiation heat loss. However, constant transport properties and a simple one-step overall reaction mechanism, which cannot consider the effects of chain-branching and the chainterminating process, were used in the paper. The experimental study carried out by Maruta et al [8] under microgravity conditions indicated that, in addition to the stretch extinction which occurred at a high stretch rate, flame extinction also occurred when the stretch rate decreased to a sufficiently low value. Numerical investigations by Guo et al [9] using the detailed chemistry and transport properties indicated that the flame extinction at a low stretch rate is due to radiation heat loss and showed a C-shaped extinction limit curve. Sung et al [10] also showed this phenomenon by numerical calculation. More recently, Ju et al [11] found the phenomenon of two stable flame branches and showed a G-shaped curve through further numerical calculations This phenomenon of two stable flame branches was also found by Buckmaster [12] using a simple asymptotic strategy. Guo et al [13] extended the investigations by considering radiation reabsorption, and found that reabsorption has little influence on the extinction limits of CH 4 air counterflow premixed flames. Furthermore, Ju et al [14] investigated the effect of the Lewis number on the extinction characteristics of stretched premixed flames by using a simple one-step reaction scheme and constant transport properties. The results show that a variation of the Lewis number causes the extinction curve to change. Most of the above studies were conducted for CH 4 air flames. Another typical fuel in practical applications is hydrogen, which has a much lower Lewis number than that of a methane air mixture. The radiating system of a H 2 air flame, where only H 2 O is the radiating species, is different from that of a CH 4 air flame, where radiating species include CO 2,H 2 O, CO and CH 4. Therefore, we can expect that the phenomenon of extinction of H 2 air counterflow premixed flames may be different from that of CH 4 air flames. In the present study, numerical investigations of counterflow premixed H 2 air flames were conducted. Firstly, attention was paid to whether the radiation extinction limit exists for stretched H 2 air premixed flames. Secondly, we tried to find the differences between the extinction phenomena of CH 4 air flames and H 2 air flames due to the change of Lewis number and radiating system. Finally, we focused our attention on the relation between the limits of counterflow premixed flames and the concentration limit of one-dimensional planar premixed flames by conducting calculations in both configurations using the same chemistry, transport properties and radiation model. 2. Numerical model We considered a flame produced by two counterflow H 2 air mixture streams which emerge from two coaxial jets. A chemically reacting boundary layer is established in the neighbourhood of the stagnation point produced by these two flows. The calculation for counterflow premixed flames assumes laminar stagnation point flow, and the boundary layer approximation is applied. The governing equations for mass, momentum, chemical species and energy equations are the same as those of our previous calculations [9, 11]. For the sake of brevity, only the energy equation is given here: ρc p dt dt dt + ρvc p dy = d ( λ dt ) dy dy KK KK h k ω k M k k=1 k=1 ρy k c pk V k dt dy + q r. (1)

Extinction of counterflow premixed H 2 air flames 461 In equation (1), q r is the sink term due to the thermal radiation, which will be described later, and KK is the number of species. Potential flow boundary conditions were used. A 19-reaction, nine-species chemical mechanism, as given in [15], was used for the H 2 air mixtures. The pressure was 1 atm, and the ambient temperature T 0 = 300 K. Thermodynamic properties were calculated from the CHEMKIN database [16], and the transport properties were calculated by using the code developed by Kee et al [17]. Our calculations were carried out basically by using the method developed by Smooke et al [18 20]. Windward differencing was used for the convective term, and the calculation included adaptive mesh refinement. An improved arc-length continuation method [11, 20, 21] was used to obtain the extinction limits. Thermal diffusion of H and H 2 was taken into account. Our previous investigation [13] using an improved radiation model, in which the radiation reabsorption is included, showed that the reabsorption has little influence on the extinction limits of counterflow premixed flames. The maximum optical thickness in all our present calculations is much less than 1.0. Therefore, the optically thin radiation model was used in the present calculation to save computation time. Thus the sink term due to the radiation heat loss, q r, may be calculated as [22] q r = 4σk p (T 4 T0 4 ) (2) where σ is the Stefan Boltzmann constant, k p is the Planck mean absorption coefficient, T 0 is the ambient temperature and T is the instantaneous local temperature. The Planck mean absorption coefficient accounts for the absorption and emission from H 2 O, which is the only dominant radiating species in H 2 air flames, and is expressed as k p = px H2 Ok p,h2 O (3) where p is pressure (atm), k p,h2 O and x H2 O denote the mean absorption coefficient and mole fraction, respectively, of H 2 O. The quantity k p,h2 O was obtained by fitting the data given in [23]. 3. Results and discussions 3.1. Flame bifurcation 3.1.1. Lower equivalence ratio flames (φ <0.1875). Figure 1 shows the variations of the maximum flame temperature versus stretch rate for three lower equivalence ratio flames (the equivalence ratios are 0.1426, 0.1466 and 0.1601, respectively). It can be found that the maximum flame temperature profile of every flame is a closed O-shaped curve. We know that the upper branch of the curve is the physically stable solution, and the lower branch is the physically unstable solution. As the stretch rate increases from a middle value, the maximum flame temperature first rises to a maximum value, then decreases and finally the flame extinguishes at a higher stretch rate. This phenomenon is well known. First, the Lewis number, defined as the ratio of the mixture thermal diffusivity to the fuel mass diffusivity of a lean H 2 air mixture, is about 0.3, i.e. significantly less than 1.0. Therefore, with increasing stretch rate, the energy gain of the reaction zone due to the preferential diffusion of the deficient species (H 2 ) becomes stronger, and thus the flame temperature rises. Meanwhile, the radiative heat loss decreases due to flame thinning, which also makes the flame temperature rise as the stretch rate increases. Both of these two factors lead to rising flame temperature as the stretch rate increases. However, with a further increase of the stretch rate, the chemical reaction cannot be completed because of the shortened residence time, and thus the flame temperature decreases and finally the flame extinguishes. Therefore, the extinction limit for each flame in

462 H Guo et al Figure 1. Maximum flame temperatures versus the stretch rate, the equivalence ratios are 0.1426, 0.1466 and 0.1601, respectively. the right-hand portion of the curve in figure 1 is caused by the excessive stretch. We call this extinction limit the stretch extinction limit. As the stretch rate decreases from a middle value, the flame temperature decreases and finally the flame also extinguishes at a low stretch rate for each equivalence ratio. This phenomenon means that not only an excessive stretch but also too low a stretch causes the flame to extinguish. The maximum flame temperature is always located at the stagnation plane regardless of the stretch rate for all lower equivalence ratio flames in figure 1. There are two factors that make the flame temperature reduce as the stretch rate decreases. First, as indicated above, the Lewis number effect weakens and thus the flame temperature reduces when the stretch rate decreases. Secondly, the radiation heat loss increases and thus the flame temperature reduces as the stretch rate decreases. However, the Lewis number effect cannot allow the flame to extinguish at a low stretch rate. This can be confirmed by conducting the calculation for an adiabatic flame (neglecting radiation heat loss), as shown in figure 2. We can find that the adiabatic flame does not extinguish at a low stretch rate, because the Lewis number effect becomes very weak at low stretch rates. The flame extinction at a low stretch rate occurs only after the radiation heat loss is considered. Therefore, the radiation heat loss is the direct factor which results in the flame extinction as the stretch rate decreases to a sufficiently low value. As for CH 4 air flames [9], we call this flame extinction limit at the low stretch rate the radiation extinction limit, to distinguish it from the stretch extinction limit. Therefore, the radiation extinction limit still exists for the H 2 air counterflow premixed flame, although its radiating system and Lewis number are different from those of a CH 4 air flame. As we analysed for CH 4 air flames [9], the mechanism of radiation extinction is as follows. We know that the heat release rate, which is proportional to the mass consumption rate of fuel, reduces because of the decreasing mass flow rate from the burner as the stretch rate decreases. On the other hand, the flame thickens and thus the radiation heat loss, which is proportional to the volume fraction of emitting species (H 2 O), increases relatively as the stretch rate decreases.

Extinction of counterflow premixed H 2 air flames 463 Figure 2. Maximum flame temperatures versus the stretch rate, the equivalence ratio is 0.1466. Figure 3. Radiation fraction and maximum flame temperature versus the stretch rate, the equivalence ratio is 0.1601. Therefore, the radiation fraction, defined as in our previous paper [9], rises rapidly, and the flame temperature drops more rapidly than the adiabatic flame as the stretch rate decreases, as shown in figures 2 and 3. This flame temperature drop causes the burning velocity to reduce rapidly. Thus the flame cannot move away from the stagnation plane when the stretch rate decreases, although the flow speed of the unburnt mixture also decreases. As the stretch rate decreases to a sufficiently low value, the flame temperature drops to such a low value that the burning velocity becomes very low and the chemical reaction cannot be completed within finite residence time. Then radiation extinction occurs. So we can conclude that the

464 H Guo et al Figure 4. Half the flame separation distance and maximum flame temperature versus the stretch rate, the equivalence ratio is 0.244. Point a is the radiation extinction limit, b is the point where twin flames merge again, c is the point where the maximum flame temperature begins to rise with decreasing stretch rate, d is the point where the distance between twin flames takes the maximum value, e is the point where twin flames begin to form, f is the point where the maximum flame temperature takes the maximum value as the stretch rate changes and g is the stretch extinction limit. direct fact which results in the radiation extinction is the radiation heat loss, being different from the stretch extinction limit caused by the excessive stretch, although they share the same phenomenon incomplete combustion. We can also note that the maximum flame temperature at the radiation extinction limit is lower than that at the stretch extinction limit for each equivalence ratio flame. This is attributed to the different residence time at these two limits. The flame is very thin due to the high stretch rate, so the residence time is very short near the stretch extinction limit. Since the flame thickens as the stretch rate decreases, the residence time is longer near the radiation extinction limit than that near the stretch extinction limit. Therefore, a flame of lower stretch rate can be sustained at a lower temperature than that of higher stretch rate, and the maximum flame temperature at the radiation extinction limit is lower than that at the stretch extinction limit. This extinction phenomenon of H 2 air flames in a lower equivalence ratio region is similar to that of CH 4 air flames [9, 11]. 3.1.2. Middle equivalence ratio flames (0.1875 φ 0.2734). With increasing equivalence ratio, the closed O-shaped curve of maximum flame temperature profile shown in figure 1 is distorted. Figure 4 shows the maximum flame temperature profile for the flame of equivalence ratio 0.244. It can be found that as the stretch rate increases from a middle value, the phenomenon is similar to those in figure 1. However, the left-hand portion of this curve is a little different from those of lower equivalence ratio flames. As the stretch rate decreases from a middle value, the maximum flame temperature first also drops, but then rises when the stretch rate further decreases from point c until point b is reached. Finally, the maximum flame temperature drops again from point b and radiation extinction occurs at point a. This difference of the maximum flame temperature profiles between lower and middle

Extinction of counterflow premixed H 2 air flames 465 equivalence ratio flames is due to the rise of the burning velocity with increasing equivalence ratio. When the stretch rate decreases from point f, the maximum flame temperature drops because of the Lewis number effect and increasing radiation heat loss (due to flame thickening). However, the burning velocities become higher for middle equivalence ratio flames than those of lower equivalence ratio flames. Therefore, the flame propagation speed begins to exceed the unburnt mixture flow speed from point e, and the flame begins to move toward the burner exit and twin flames are formed. Then the distance, based on the maximum flame temperature, between twin flames increases with decreasing stretch rate, as shown in figure 4 (half the flame separation distance is shown in the figure), until point d, where the distance between twin flames takes the maximum value, is reached. In the meantime, there are two factors that result in a more rapid drop of the maximum flame temperature than that of the adiabatic flame. First, like lower equivalence ratio flames, the flame thickens as the stretch rate decreases and thus the direct radiation heat loss from flame to environment increases. Secondly, a temperature dip between twin flames is formed because of the radiation heat loss, and thus there are conductive heat losses from the reaction zone to the stagnation plane. It should be noted that these conductive heat losses are also caused, but indirectly, by radiation heat loss. Both the direct radiation heat loss and the conductive heat losses accelerate the flame temperature drop as the stretch rate decreases. As a result, the accelerated flame temperature drop allows the burning velocity to reduce more rapidly. When the stretch rate decreases to point d, the flame propagation speed becomes lower than the local flow velocity of the unburnt mixture upstream of the flame front. Therefore, the twin flames move toward the stagnation plane again, and the twin flame separation distance (based on the maximum flame temperature) reduces with the further decrease of the stretch rate. When twin flames are near the stagnation plane (at point c), the local stretch rate in the reaction zone is strengthened due to the streamline divergence, and therefore the flame temperature again rises slightly with decreasing stretch rate until point b, where twin flames merge again, is reached. Then the flame thickens and the flame temperature drops continuously with decreasing stretch rate until radiation extinction occurs at point a. Thus the distorted O-shaped temperature profile is a result of the combined effect of the Lewis number and the radiation heat loss. Please note that the maximum flame temperature rise from point c to b does not occur for the middle equivalence ratio CH 4 air flames [9, 11]. This is because the Lewis number of the CH 4 air mixture is 0.98 (near 1.0), while that of the H 2 air mixture is 0.3 (much less than 1.0). Therefore, when twin flames of CH 4 air move back to the stagnation plane due to the radiation heat loss, the effect of the Lewis number is not strong enough to allow the flame temperature to rise, while the H 2 air flame does due to its much lower Lewis number. So we can say that the transition of the O-shaped temperature profile to a distorted O-shaped profile only occurs for a mixture with a Lewis number much less than 1.0, such as a H 2 air mixture. This radiation extinction process can be clearly shown by the flame temperature and H 2 mass fraction distributions of three typical stretch rates in figure 5. The maximum flame temperature is located at the stagnation plane at a stretch rate of 100 s 1. When the stretch rate decreases to 1.0 s 1, the flame temperature is lower than that at the stretch rate of 100 s 1 and a temperature dip between the twin flames is formed due to radiation heat loss. The twin flames merge again and incomplete combustion occurs as the stretch rate decreases further to 0.16 s 1, which is near the radiation extinction limit. Although the radiation extinction of a stretched H 2 air flame has not been observed in experiments, a similar flame extinction process for CH 4 air was indeed observed by the microgravity experiment [8].

466 H Guo et al Figure 5. Temperature and H 2 mass fraction distributions, the equivalence ratio is 0.244. All flames with middle equivalence ratios (0.1875 φ 0.2734) have a similar extinction process, as shown in figure 6. 3.1.3. Higher equivalence ratio flames (φ 0.3033). Figure 7 shows the maximum flame temperature profiles for flames with equivalence ratios of 0.3033 and 0.2734, respectively. We can find that the right-hand portions of the two curves are similar, but the left-hand portions are different. The stable flame branch of equivalence ratio 0.3033 divides into two branches, abc and ef, being different from the flame of equivalence ratio 0.2734. Furthermore, the ef branch extends to point d, where the stretch rate is higher than that at point c. There are four turning points, a, c, d and f, along the closed curve. From the fold bifurcation theory and the previous studies [11, 12], we know that every turning point should be a separation point between the

Extinction of counterflow premixed H 2 air flames 467 Figure 6. Maximum flame temperatures versus the stretch rate, the equivalence ratios are 0.1875, 0.2155, 0.244 and 0.2734, respectively. stable solution branch and the unstable solution branch. At the same time, we also know that the branch abc is the maximum flame temperature curve of a stable flame branch like that for the flame of equivalence ratio 0.2734. Therefore, cd is a maximum flame temperature curve of the physically unstable branch, def is another maximum flame temperature curve of the physically stable flame branch and fga is another maximum flame temperature curve of the physically unstable flame branch. Because the maximum flame temperature of branch def is lower than that of branch abc, we call branch def a weak flame branch and abc a normal flame branch, as in our study of a CH 4 air counterflow premixed flame [11]. To analyse the flame bifurcation mechanism in figure 7, half the flame separation distance, based on the maximum flame temperature position, is plotted versus the stretch rate in figure 8 for the equivalence ratio 0.3033. As discussed for middle equivalence ratio flames, when the stretch rate decreases from point b, where the maximum flame temperature takes its maximum value, the flame separation distance first increases and then decreases. However, when the stretch rate decreases to point c, the maximum flame temperature of a normal flame drops vertically there, although the flame separation distance has not decreased to zero. Then there is no solution existing for the normal flame branch for a further decrease of the stretch rate. Meanwhile, the weak flame exists at stretch rates lower than that of point c. Therefore, point c can be thought of as a jump limit, where the normal flame suddenly jumps to the weak flame. This phenomenon can be explained as follows. Unlike the flame of equivalence ratio 0.2734, the flame of equivalence ratio 0.3033 cannot move back to the stagnation plane with a continuous decrease of stretch rate due to the burning velocity strengthening when the equivalence ratio is higher. However, the direct radiation heat loss increases relatively, and the conductive heat losses from the flame zone to the stagnation plane increase as well due to the widened flame separation distance and larger temperature dip, compared with the heat losses of equivalence ratio 0.2734. The increase of these two kinds of heat losses results in the flame temperature dropping vertically and the normal flame jumps to the weak flame at point c before its twin flames move back to the stagnation plane. Of course, the flame extinguishes at the stretch

468 H Guo et al Figure 7. Maximum flame temperatures versus the stretch rate, the equivalence ratios are 0.2734 and 0.3033, respectively. Point a is the stretch extinction limit of the normal flame branch, b is the point where the maximum temperature of the normal flame takes the maximum value as the stretch rate changes, c is the jump limit of the normal flame branch, d is the jump limit of the weak flame branch, e is the point where the maximum temperature of the weak flame takes the maximum value as the stretch rate changes and point f is the radiation extinction limit of the weak flame branch. extinction limit (a) as the stretch rate is increased. Therefore, there are two limits for the normal flame branch the jump limit (c) and the stretch extinction limit (a). There is no such jump limit for the normal flame branch of CH 4 air counterflow premixed flames [11]. So the appearance of this jump limit is the result of a lower Lewis number. Therefore, although the H 2 air mixture is simpler, its extinction mechanism is more complex than CH 4 air premixed flames due to its much lower Lewis number (much less than 1.0). For the weak flame branch, the radiation extinction limit (f ) is reached with decreasing stretch rate due to the flame thickening. On the other hand, with increasing stretch rate the flame temperature rises and the burning velocity increases first until point e is reached because of decreasing radiation heat loss. Then the flame moves away from the stagnation plane and the flame temperature begins to go down due to weakening of the Lewis number effect. However,

Extinction of counterflow premixed H 2 air flames 469 Figure 8. Half the flame separation distance versus the stretch rate, the equivalence ratio is 0.3033. Point a is the stretch extinction limit of the normal flame branch, b is the point where the maximum temperature of the normal flame takes the maximum value as the stretch rate changes, c is the jump limit of the normal flame branch, d is the jump limit of the weak flame branch, e is the point where the maximum temperature of the weak flame takes the maximum value as the stretch rate changes and point f is the radiation extinction limit of the weak flame branch. the drop rate of the flame temperature reduces with a further increase of the stretch rate due to decreasing radiation heat loss. Finally, the flame temperature rises again near point d and suddenly rises vertically at point d. Then there is no solution existing for the weak flame when the stretch rate is greater than that of point d. So point d can be thought of as a jump limit, where the weak flame jumps up to the normal flame branch, of the weak flame. Therefore, there are also two limits for the weak flame, the radiation extinction limit (f ) and the jump limit (d). We can conclude that there are two stable flame branches, normal and weak flame branches, for higher equivalence ratio H 2 air counterflow premixed flames. For each of these two branches, there are two limits, the stretch extinction and the jump limits for the normal flame branch, and the radiation extinction and the jump limits for the weak flame branch. When the stretch rate is within the range between points c and d, there are two solutions existing at the same stretch rate, a normal flame solution and a weak flame solution. The normal flame branch has a wider flame separation distance (based on the maximum flame temperature) and higher temperature, and the weak flame has a narrower flame separation distance and lower temperature, as shown in figures 7 and 8. Figure 9 shows the radiation fraction distribution of the 0.3033 equivalence ratio flame. The stretch rate at the jump limit of the normal flame is greater than that at the radiation extinction limit of the weak flame. However, the value of the stretch rate at the jump limit of the normal flame decreases sharply with increasing equivalence ratio, because the burning velocity increases and thus the twin flames can move further away from the stagnation plane when the stretch rate decreases. When the equivalence ratio is 0.3124, the stretch rate at the jump limit of the normal flame almost takes the same value as that at the radiation extinction limit of the weak flame, as shown in figure 10. With a further increase of the equivalence ratio, the jump limit of the normal flame disappears and the closed curve of the

470 H Guo et al Figure 9. The radiation fraction versus the stretch rate, the equivalence ratio is 0.3033. Point a is the stretch extinction limit of the normal flame branch, b is the point where the maximum temperature of the normal flame takes the maximum value as the stretch rate changes, c is the jump limit of the normal flame branch, d is the jump limit of the weak flame branch, e is the point where the maximum temperature of the weak flame takes the maximum value as the stretch rate changes and point f is the radiation extinction limit of the weak flame branch. maximum flame temperature profile opens, such as the flame of equivalence ratio 0.3185. Then the maximum flame temperature curve of the normal flame branch can be extended to zero stretch rate, at which the solution should be the solution of a one-dimensional planar premixed flame. This is because the flame burning velocity is so high that the twin flames can move continuously toward the burner exit, like two one-dimensional planar premixed flames, when the stretch rate is decreased continuously for such a higher equivalence ratio. Therefore, we can expect that the solution of a one-dimensional planar premixed flame exists only when the equivalence ratio is greater than the value above which the jump limit of the normal flame does not exist and the closed curve of the stretched flame temperature profile opens. Both the radiation extinction and jump limits of a weak flame still exist when the equivalence ratio increases further. This phenomenon is similar to that of CH 4 air flames. Although this two-flame branch phenomenon has never been observed in experiments, it was found in the numerical and theoretical studies of CH 4 air counterflow premixed flames [11, 12]. It is expected that this phenomenon can be confirmed by microgravity or zero-gravity experiments in the future. 3.2. Concentration limit of one-dimensional planar premixed flames In order to find the relation between counterflow premixed flames and one-dimensional planar premixed flames, we also conducted calculations of one-dimensional planar premixed flames by considering radiation heat loss. The chemistry and radiation model used are the same as those described above. Figure 11 shows the variations of flame temperature and burning velocity versus H 2 concentration. We find that both the flame temperature and the burning velocity drops vertically at an equivalence ratio of 0.314 (fuel percentage of 11.66), below

Extinction of counterflow premixed H 2 air flames 471 Figure 10. The maximum flame temperatures versus the stretch rate, the equivalence ratios are 0.3124 and 0.3185. which the flame cannot exist. This means that the concentration limit of a one-dimensional planar H 2 air premixed flame is 11.66% (an equivalence ratio of 0.314). This concentration limit obtained by us is much higher than that, about 4.0%, given in the textbooks [24, 25] for H 2 air premixed flames. The difference between the present result and that in the textbooks is mainly attributed to the non-planar cellular flame structure, which was not considered in our calculation, in near-limit one-dimensional H 2 air premixed flames. It has been known for over 40 years that when flames travel upward through the mixtures containing from 4.1% 10% hydrogen, a portion of the hydrogen remains unburnt, while the hydrogen air mixture must be enriched to over 10% hydrogen in order to support a flame in downward flame propagation [25]. This phenomenon is due to the appearance of discrete curved flames with a non-planar cellular structure, which appears to be rather resilient to heat losses and may survive far beyond the extinction point of the associated planar flame [26], in the near-lean-limit H 2 air premixed flame. No planar premixed flames can be found below 10% hydrogen. Therefore, the present calculation gives a numerical result, which is very close to the experimental result, of the flammability limit of one-

472 H Guo et al Figure 11. Flame temperature and burning velocity of a one-dimensional premixed flame versus H 2 concentration. dimensional planar premixed H 2 air flames. Although the limit listed in the textbooks [24, 25] is much lower than the present numerical result, that is the flammable region of cellular flame. This result shows that a cellular H 2 air flame can survive under the flammability limit of a planar premixed flame. Actually, the physics of cellular structure is essentially similar to that of a stretched counterflow premixed flame and steady flame balls have been shown to exist in the sub-limit condition [27]. Therefore, counterflow H 2 air premixed flames can exist under the flammability limit of a one-dimensional planar premixed flame, as will be discussed in the next section. 3.3. Flammable regions The flammable regions of normal and weak flames can be formed by plotting all limits on the a φ-plane, as shown in figure 12. ABC is a C-shaped curve, which is similar to that we obtained for the stretched CH 4 air premixed flames [9]. AB is the stretch extinction limit branch, BC is the radiation extinction limit branch and DE is the jump limit curve of the normal flame. DG is the jump limit curve of a weak flame and CI is the radiation extinction limit curve of a weak flame. The equivalence ratio value at point K is the concentration limit of a one-dimensional planar premixed flame. We can find that point B, where the stretch extinction limit branch and the radiation extinction limit branch merge, is the lowest fuel concentration limit, below which the counterflow premixed H 2 air flames cannot exist regardless of the stretch rate. Neither the stretch extinction limit branch (AB) nor the radiation extinction limit branch (BC) of a normal flame can be extended to the concentration limit of a one-dimensional planar premixed flame. Only the maximum equivalence ratio value (point E), above which the jump limit of a normal flame does not exist, is the concentration limit of a one-dimensional planar premixed flame. At this equivalence ratio, the stretch rates of the normal flame jump limit and the weak flame radiation extinction limit are the same.

Extinction of counterflow premixed H 2 air flames 473 Figure 12. Flammable regions of H 2 air counterflow premixed flames, where point K is the flammability limit of a one-dimensional premixed flame. Figure 13. Flammable regions of CH 4 air counterflow premixed flames [11], where point E is the flammability limit of a one-dimensional planar premixed CH 4 air flame. The velocity gradient corresponds to the stretch rate. The right-hand side of ABCDEK is the flammable region of a normal flame, and the weak flame can exist within the region GDCI. GDEI is a region within which both normal and

474 H Guo et al weak flames can exist. The stretch rates at both the radiation and jump limits of a weak flame decrease with the rise of the equivalence ratio first, and then slightly increase with the further rise of the equivalence ratio from an equivalence ratio of about 0.794. The flammable regions of normal and weak flames of CH 4 air counterflow premixed flames [11] are given in figure 13 for comparison with the present results. We can find that for ah 2 air mixture, the difference between the flammability of a counterflow premixed flame and the flammability limit of a one-dimensional planar premixed flame is greater than that for ach 4 air mixture. This means that the extension of the flammability limit due to the stretch is amplified for a H 2 air flame. It is due to decreasing Lewis number. Therefore, we can conclude that the Lewis number has a dramatic effect on the extension of the flammable region caused by the flame stretch. 4. Conclusions Numerical calculations of the counterflow premixed H 2 air flames have been conducted. Emphasis has been placed on the effect of radiation heat loss and the mechanisms of flame bifurcation and the effect of the Lewis number. Also attention was paid to the relation between the counterflow premixed flames and the one-dimensional planar premixed flame. Radiation heat loss has an important impact on the flames at the lower stretch rate region. When the equivalence ratio is lower (φ 0.1875), the temperature profile of every flame is a closed O-shaped curve, which is similar to those of stretched CH 4 air premixed flames at a lower equivalence ratio region. There are two extinction limits for every flame, the stretch extinction limit induced by the excessive stretch at a higher stretch rate and the radiation extinction limit induced by the radiation heat loss at a lower stretch rate. For middle equivalence ratio (0.1875 φ 0.2734) flames, the temperature profile of every flame becomes a distorted O-shaped curve due to the combined effect of the radiation heat loss and the Lewis number. This phenomenon does not appear for CH 4 air counterflow premixed flames. There are also two extinction limits for every flame, the stretch extinction limit and the radiation extinction limit. However, the temperature profiles of higher equivalence ratio (φ 0.3033) flames are different from those of lower and middle equivalence ratio flames. First, when the equivalence ratio is lower than a critical value (0.314), the temperature profile of every equivalence ratio is still a closed curve, but there are four limits along the curve. The curve is divided into two stable flame branches and two unstable branches by these four limits. For every stable flame branch, there are two limits, a stretch extinction limit and a jump limit for the normal flame branch and a radiation extinction limit and a jump limit for the weak flame branch. This jump limit of the normal flame branch does not exist for the CH 4 air mixture. It is the result of a lower Lewis number. If the equivalence ratio is greater than this critical value (0.314), the jump limit of a normal flame does not exist, and the closed flame temperature profile of every equivalence ratio opens. Then the normal flame branch can be extended to zero stretch rate, at which the solution of the counterflow flame is the solution of a one-dimensional planar premixed flame. However, the two limits of a weak flame still exist. The results show that this critical equivalence ratio for the opening up of the temperature profile corresponds to the lean limit of the one-dimensional planar premixed H 2 air flame. The comparison between the extinction curves of CH 4 air and H 2 air flames shows that the extension of the flammability limit of stretched flames is amplified by the decrease of the Lewis number.

Extinction of counterflow premixed H 2 air flames 475 References [1] Spalding D B 1957 Proc. Soc. London A 240 83 100 [2] Lakshmisha K N, Paul P J and Mukunda H S 1990 23rd Int. Symp. on Combustion (Pittsburgh, PA: Combustion Institute) pp 433 40 [3] Tsuji H and Yamaoka I 1982 19th Int. Symp. on Combustion (Pittsburgh, PA: Combustion Institute) p 1533 [4] Sato J 1982 19th Int. Symp. on Combustion (Pittsburgh, PA: Combustion Institute) p 1541 [5] Sohrab S H and Law C K 1984 Int. J. Heat Mass Transfer 27 291 [6] Egolfopoulos F N 1994 25th Int. Symp. on Combustion (Pittsburgh, PA: Combustion Institute) p 1375 [7] Platt J A and T ien J S 1990 Chemical and Physical Process in Combustion (1990 Fall Technical Meeting, Eastern Section of the Combustion Institute) p 31-1 [8] Maruta K, Yoshida M, Ju Y and Niioka T 1996 26th Int. Symp. on Combustion (Pittsburgh, PA: Combustion Institute) p 1283 [9] Guo H, Ju Y, Maruta K, Niioka T and Liu F 1997 Combust. Flame 109 639 46 [10] Sung C J and Law C K 1996 26th Int. Symp. on Combustion (Pittsburgh, PA: Combustion Institute) pp 865 73 [11] Ju Y, Guo H, Maruta K and Liu F 1997 J. Fluid Mechanics 342 315 [12] Buckmaster J 1997 Combust. Theory Modelling 1 1 11 [13] Guo H, Ju Y, Maruta K, Niioka T and Liu F 1998 Combust. Sci. Technol. 135 49 64 [14] Ju Y, Guo H, Maruta K and Niioka T 1998 Combust. Flame 113 603 14 [15] Buckmaster J, Smooke M and Giovangigli V 1993 Combust. Flame 94 113 24 [16] Kee R J, Grcar J F, Smooke M D and Miller J A 1994 SANDIA report SAND85-8240 [17] Kee R J, Warnatz J and Miller J A 1983 SANDIA report SAND 83-8209 [18] Smooke M D 1982 J. Comput. Phys. 48 72 [19] Giovangigli V and Smooke M D 1989 Appl. Numer. Math. 5 305 [20] Giovangigli V and Smooke M D 1987 Combust. Sci. Technol. 53 23 [21] Kee R J, Miller J A, Evans G H and Dixon-Lewis G 1988 22nd Int. Symp. on Combustion (Pittsburgh, PA: Combustion Institute) pp 1479 [22] Modest M F 1993 Radiative Heat Transfer (New York: McGraw-Hill) [23] Tien C L 1967 Adv. Heat Transfer 5 253 [24] Kuo K K 1986 Principles of Combustion (New York: Wiley) p 278 [25] Lewis B and von Elbe G 1961 Combustion and Explosions of Gases (New York: Academic) pp 315 23 [26] Kagan L and Sivashinsky G 1997 Combust. Flame 108 220 [27] Ronney P 1990 Combust. Flame 82 1 14