Sie sind auf Seite 1von 9

Seik Mansoor Ali

Safety Research Institute, Atomic Energy Regulatory Board, Kalpakkam-603 102, India

A Numerical Study of Concurrent Flame Propagation Over Methanol Pool Surface


Concurrent ame spread over methanol pool surface under atmospheric conditions and normal gravity has been numerically investigated using a transient, two-phase, reacting ow model. The average ame spread velocities for different concurrent air velocities predicted using the model are quite close to the experimental data available in the literature. As the air velocity is increased, the fuel consumption rate increases and aids in faster ame spread process. The ame initially anchors around the leading edge of the pool and the ame tip spreads over the pool surface. The rate of propagation of ame tip along the surface is seen to be steady without uctuations. The ame spread velocity is found to be nonuniform as the ame spreads along the pool surface. The ame spread velocity is seen to be higher initially. It then decreases up to a point when the ame has propagated to around 40% to 50% of the pool length. At this position, a secondary ame anchoring point is observed, which propagates toward the trailing edge of the pool. As a result, there is an increasing trend observed in the ame spread velocity. As the air velocity is increased, the initial ame anchoring point moves downstream of the leading edge of the fuel pool. The variations of interface quantities depend on the initial ame anchoring location and the attainment of thermodynamic equilibrium between the liquid- and gas-phases. [DOI: 10.1115/1.4005111] Keywords: concurrent ame spread, methanol pool, liquidgas interface, ame spread velocity, ame anchoring point

Vasudevan Raghavan1
e-mail: raghavan@iitm.ac.in Department of Mechanical Engineering, Indian Institute of Technology Madras, Chennai-600 036, India

K. Velusamy
Indira Gandhi Center for Atomic Research, Kalpakkam-603 102, India

Shaligram Tiwari
Department of Mechanical Engineering, Indian Institute of Technology Madras, Chennai-600 036, India

Introduction

Flame propagation over spilled combustible liquid fuels is a serious re safety issue and presents a practical re safety problem. Over the years, a number of theoretical, experimental, and numerical studies have appeared in literature on ame spread over condensed fuel surfaces. Early theories [1] on ame spread over liquid fuels reported the development of uid motion below the propagating ame when fuel is at sub ash temperatures. The concept of ame spread on liquid pool surfaces and the various mechanisms controlling it have been discussed by many researchers [25]. The role of liquid fuel temperature on the ame spread process has also been elucidated in some of these early studies. Flame spread phenomena were classied into four groups based on the initial temperature of liquid [6]. According to this classication, in the subash regime, the spread could be pseudouniform or pulsating in nature, whereas in the near-ash and super-ash regime, the ame propagation would be uniform. The pulsating behavior was shown to be a consequence of ame interaction with surface ow, which leads to abrupt periodic acceleration of the ame front [6,7]. Excellent review articles on ame spread literature and discussions on the predominant mechanisms of heat transfer that control the ame spread rate in each regime are available [8,9]. Model problems demonstrating integral analysis on ame spread under wind aided conditions have also been reported [10]. Based on detailed experimental investigations, the mechanism of ame propagation over methanol surfaces in super-ash and near-ash regime, under concurrent as well as opposed air ow conditions were elucidated [11,12]. Later, experimental investigation on ame spread over porous solids soaked with kerosene
Corresponding author. Contributed by the Heat Transfer Division of ASME for publication in the JOURNAL OF HEAT TRANSFER. Manuscript received April 20, 2011; nal manuscript received September 12, 2011; published online February 15, 2012. Assoc. Editor: Ali Ebadian.
1

brought out the role played by behavior of fuel in the mixed media. The two-phase heat transfer in controlling the ame spread process was explained [13,14]. Different experimental techniques [1517] have been applied to investigate the pulsating ame spread across various liquid fuels. Experimental studies have been conducted on ame spread over n-butanol pool under normal as well as microgravity environment and the effect of low speed concurrent and opposed air ow on the ame spread characteristics have been brought out [18]. Recently, experimental studies were reported on the effect of oxygen on the ame spread over several combustible liquids [19]. Over the years, various numerical models to study the structure of subsurface liquid ows preceding a ame, spreading at a steady rate, have been reported [20,21]. The effect of Reynolds number, Prandtl number and the liquid surface tension was investigated to elucidate the role played by the hydrodynamics of the liquid layer in controlling the ame spread process. Some of these early models were concerned with liquid-phase only and neglected the gas liquid phase coupling that is essential to describe the ame spread process. An effective conductivity model for ame spread over shallow sub ash liquid pool layer has been presented [22]. A two-dimensional numerical model that coupled both the phases to capture the oscillatory ame spread over a liquid (methanol) at sub ash initial temperatures and employing a nite-rate chemical kinetics model has also been reported [23,24]. In these works, the importance of surface tension induced liquid-phase convective motion was investigated by varying the viscosity over a wide range. A detailed, transient two-dimensional numerical model to provide an accurate description of ame spread over liquid pools in both pulsating and uniform regimes was presented later [25]. This model could correctly predict the instantaneous ame speed in steady and pulsating modes. Besides, it could also predict the pool temperature at which transition from pulsating to steady ame spread occurs. Numerical studies to examine the inuence of forced opposed air ow conditions on ame spread over n-propanol pools are available [26,27]. Flame propagation at APRIL 2012, Vol. 134 / 041202-1

Journal of Heat Transfer

C 2012 by ASME Copyright V

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/12/2013 Terms of Use: http://asme.org/terms

normal as well as zero gravity conditions has been discussed. Similar numerical studies were also reported [28] on axisymmetric ame propagation over propanol pools in quiescent (no ow) conditions under both normal and zero gravity. Based on the above literature survey, it appears that numerical studies on ame propagation over a liquid fuel maintained above its ash point temperature and subjected to a concurrent air ow are rare. The probable reason for this could be that, in such cases, the propagation mechanism over the fuel surface is directly inuenced by the presence of sufcient amount of fuel vapor that forms a combustible mixture, especially for volatile fuels such as methanol. However, it is known that the ame spread rate for these conditions could be very fast depending on the concurrent air velocity and hence could be complex. Furthermore, in case of ammable liquid spills exposed to such conditions, especially those of alcohols such as methanol and ethanol, this could lead to a serious re safety issue. The prediction of ame spread rate under concurrent air ow conditions, along with the details of ow and thermal elds, forms the motivation for the present study.

Gas-phase:   @ @ @ @ @ Yi qYi quYi qvYi qDi;m @t @x @y @x @x   @ @ Yi _i qDi;m x @y @y

(2)

_ i denotes the net mass based rate of production of In Eq. (2), x species i per unit volume and Di,m is the mass diffusivity of any species diffusing into the mixture. For liquid-phase, only fuel conservation is solved in the binary system. In the absence of chemical reaction in liquid-phase, the species conservation is written as Liquid-phase:  ! @ @ Yf @ @ @ qYf quYf qvYf qD12 @t @x @y @x @x  ! @ Yf @ qD12 @y @y

(3)

Description of Numerical Model

The following assumptions are made in formulating the mathematical model for transient ame propagation on fuel pool surfaces: 1. Flow is two-dimensional and laminar. 2. Fuel undergoes complete combustion through a global singlestep reaction mechanism. However, dissociation of CO2 is also considered through a partial equilibrium approach. 3. Radiative heat addition to the liquid fuel pool is negligible when compared to the conductive heat addition. 4. Properties are functions of temperature and composition only. 5. Dufour effect and species diffusion due to pressure are negligible in the gas and liquid-phase. Soret effect is also assumed to be negligible. 6. Liquid surface regression is slow enough to be neglected. 7. Gasliquid (interface) remains planar throughout the transients. In this formulation, since the fuel is an alcohol, the liquid-phase has two species including water in addition to the methanol. Therefore, solutal as well as thermal Marangoni convection effects have been modeled at the gasliquid interface. The governing differential equations in gas- and liquid-phases for twodimensional, transient, reacting ow simulations are provided below. The initial and boundary conditions are also presented subsequently.

In Eq. (3), D12 represents the binary diffusion coefcient in liquid-phase and subscript f represents fuel. (3) x-Momentum equation (for both gas-phase and liquidphase) @ @ @ @ rxx @ syx qu quu quv @t @x @y @x @y (4)

(4) y-Momentum equation Gas phase: This includes a term to simulate buoyancy induced ow due to density differenceq1 qg, where the subscript 1 represents the ambient conditions and g is acceleration due to gravity @ @ @ @ @ qv quv qvv sxy ryy @t @x @y @x @y q1 qg Liquid-phase: @ @ @ 2 @ sxy @ ryy qv quv qv @t @x @y @x @y (6)

(5)

The stress terms, r and s, in Eqs. (5) and (6), can be written using viscosity (l) as @u @v rxx p 2l ; ryy p 2l ; @x @y   @u @v sxy syx l @y @x

2.1 Governing Equations. The governing equations for mass, momentum, species, and energy conservation in the gasand the liquid-phases are given in conservative differential form in Cartesian coordinates (x, y), in terms of variables such as velocities (u, v), pressure (p), density (q), mass fraction of species i (Yi), and temperature (T). (1) Continuity equation (for both gas-phase and liquid-phase) @q @ @ qu qv 0 @t @x @y

(7)

(1)

(2) Species conservation equations There are six species in the gas-phase: CH3OH and O2 (reactants), CO, CO2, and H2O vapor (products) and N2 (an inert species). Two species, CH3OH (liquid) and H2O are considered in the liquid-phase. For any particular species i, the conservation equations in gas- and liquid-phases are of the following form: 041202-2 / Vol. 134, APRIL 2012

(5) Energy equation The energy conservation equation for gas-phase includes enthalpy transport by species, heat generation due to chemical reaction and radiative heat ux. In the liquid-phase, only enthalpy transport by species is considered in the present formulation. The governing equations for gas- and liquid-phases are given below. Gas-phase:     @ @ @ @ @T @ @T qCp T quCp T qvCp T k k @t @x @y @x @x @y @y   n n X X @ @ Y i _ 000 _ i Dhf ;i q qDi;m Cp;i T x R @x @x i1 i1  ! @ @ Yi qDi;m Cp;i T (8) @y @y Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/12/2013 Terms of Use: http://asme.org/terms

Liquid-phase:   @ @ @ @ @T qCp T quCp T qvCp T k @t @x @y @x @x   X   2 @ Yj @ @T @ k qD12 Cp;j T @y @y @x @x j1  ! @ Yj @ qD12 Cp;j T @y @y

Initial conditions for liquid-phase (subscript 0): T0 T1; u v 0; Yf0 1 Gas-phase boundary conditions: The following conditions exist at the gas-phase boundaries (Fig. 1): (1) Inlet Atmospheric air with uniform inlet velocity u1 and temperature T1 enters the domain. The mass fractions of oxygen and nitrogen correspond to those of ambient air. Hence, at the gasphase domain inlet, u u1, v 0, T T1, YO2 0.23, and YN2 0.77. (2) Wall At the walls exposed to the gas-phase, located before and after the fuel pool surface, no-slip and impermeability boundary conditions are prescribed for velocities, while temperature and species mass fractions are prescribed zero diffusive ux values. (3) Top A free boundary condition is employed at the top of the domain. The height of the gas-phase domain is selected in such a way that this free boundary condition can be implemented. All the solution variables are extrapolated in the direction normal to this boundary from the interior nodes using three point polynomial ts using ^). In the case of any zero diffusive ux in the normal direction (n reverse ow (v < 0) at this boundary, the temperature is set as 300 K and mass fractions of oxygen and nitrogen, are set as 0.23 and 0.77, respectively. These conditions can be represented as follows:
@/ @n ^ 0, where / u; v; T ; Yi , for YN2 0.77, and p p1 for v < 0

(9)

The quantity Dhf ;i in Eq. (8) is the enthalpy of formation for ith species, k represents mixture thermal conductivity, Cp represents mixture specic heat, and Cp,i represents specic heat of ith spe_ 000 cies. The radiative heat lossq R is calculated assuming the optically thin approximation with emission-only formulation as presented in Ali et al. [29]. The overall reaction rate in gas-phase has been calculated assuming a single-step global reaction mechanism for methanol-air oxidation and considering only the dissociation of CO2 through a partial equilibrium approach. A detailed discussion of the methodology used to obtain the overall reaction rate is provided in Ali et al. [29]. 2.2 Computational Domain. For the present two-dimensional conguration, the methanol pool is taken to be 100 cm long and 3 cm deep. Thin at plates having 25 cm length are assumed to be placed in ush with the top surface of the fuel pool at its leading as well as the trailing edges. Initially, the fuel pool consists of pure methanol. The computational domain in the present problem consists of a liquid-phase region of 100 cm 3 cm and a gas phase region of 150 cm 150 cm. The fuel is ignited at the leading edge by placing an ignition source at a height of 0.05 cm from its surface. The size of the ignition source is 0.25 cm 0.05 cm, which is small enough to provide a localized volumetric heat source close to the fuel surface. The free stream air ow over the fuel surface is from left to right in all the cases. The direction of gravity vector is in the perpendicular direction to the liquid pool surface. The computational domain is shown in Fig. 1. 2.3 Initial and Boundary Conditions. The initial ow elds within the gas- and liquid-phases are assumed to be stationary. Both air and fuel are present at the ambient temperature of 300 K initially. Therefore, Ambient conditions: p1 101325 N/m2; T1 300 K; YO2 0.23, YN2 0.77 Initial conditions for gas-phase: T0 T1; p0 p1; u v 0; YO2 0.23; YN2 0.77

v > 0 and T T1, YO2 0.23,

(4) Exit Boundary conditions, similar to the one applied at the top boundary, are employed here. These are given below
@/ ^ 0, where, / u; v; T ; Yi , for u > 0 T T1, YO2 0.23, @n YN2 0.77, and p p1 for u < 0

Liquid-phase: The following conditions exist at the gas-phase boundaries: (1) Container walls No-slip boundary conditions are prescribed for velocities. For temperature and mass fractions a zero diffusive ux in the normal ^) is prescribed. Therefore direction (n u 0; v 0; @ T @ Yi 0 ^ @n ^ @n

Pool interface conditions In order to solve the governing equations, the values of temperature and species mass fractions in the gas-phase and the velocity components in both gas and liquid-phase are required at the interface. The relevant governing equations for the interface are as follows: (1) Conservation of mass: The mass ux at any location on _ 00 pool surface (m x ) is given as: _ 00 m x qs vs g qs vs l (10)

where the subscript s represents interface and g, l represent the gas- and liquid-phases, respectively. (2) Continuity of tangential velocity ug;s ul;s (11)

Fig. 1

Computational domain

(3) Continuity of shear stress: This includes the gradient of surface tension (rs ) along x-direction. Further, the surface tension is evaluated as a function of temperature and concentrations of fuel and water at the interface. APRIL 2012, Vol. 134 / 041202-3

Journal of Heat Transfer

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/12/2013 Terms of Use: http://asme.org/terms

lg;s

@u @v @y @x

! ll;s
g ;s

@u @v @y @x

 @ rs    l;s @ x s

(12)

(4) Conservation of species  !  ! @ Yf @ Yf 00 _ 00 _ Fuel : m Y q D m Y q D f ;m 12 s s x f ;s x f ;s @y s g @y s l ! @ Yw _ 00 Water : m x Yw;s qs Dw;m @y s g  ! @ Yw 00 _ x Yw;s qs D12 m @y s l   @ Yi @y ! 0


s g

(13a)

(13b)

Fig. 3 Variation of ame spread velocity along the pool surface

_ 00 Other species : m x Yi;s qs Di;m

(13c) grid system is that pressure need not be specied at the walls in either the liquid- or the gas-phase regions. 3.2 Discretization Method and Solution Procedure. An inhouse numerical code is used for solving the governing conservation equations. The gas and liquid-phase governing equations are discretized using the conventional nite volume method on an Eulerian domain, using the partially staggered grid arrangement as described earlier. The nite volume integration of these equations over space and time generates a set of algebraic equations for the conservation of relevant dependent variables. Solution of these algebraic equations ensures that these dependent variables satisfy the integral form of the conservation equation for each control volume and also for the entire computational domain. The numerical method uses upwind scheme for convective terms and SIMPLE [30] algorithm for the pressure velocity coupling. The mass fraction conservation equations and the energy equation are solved using a smaller time step value for a certain number of inner iterations due to the nonlinear source terms present in those equations. All the thermophysical properties are updated as a function of temperature using appropriate correlations. The transient explicit marching procedure is carried out until either a steady-state or a time independent oscillatory solution is obtained.

(5) Conservation of energy " !  # 2 2 X X @ Yj dT   _ 00 m Y L q D L k 12 j s s x j;s j dy s g @y j1 j1 l ! dT  ks  0 dy s l

(14)

where Lj is the latent heat of vaporization of jth species. (6) Phase equilibrium: Using the activity coefcient c the phase equilibrium is represented as Xi;g ci Xi;l pi p (15)

Numerical Method

3.1 Grid system. The physical domain consists of gas-phase and liquid-phase regions as described earlier. A partially staggered grid arrangement with quadrilateral cells is employed for the numerical solution. Two separate types of control volumes are used; variables like u, v, T, and Yi are all evaluated at the centroid of a control volume. Only pressure is offset and is evaluated separately in another control volume. Both these control volumes require separate calculations for volume, area of control volume faces as well as an appropriate interpolation scheme for uxes across the control volume faces. An attractive feature of the above

Results and Discussion

Concurrent ame spread over a 1 m long methanol pool, under atmospheric conditions and normal gravity is analyzed systematically. Free stream air ow parallel to the fuel surface is varied from 1.3 m/s to 5.1 m/s. The corresponding free stream Reynolds

Fig. 2 Temporal variations of (a) the ame tip location and (b) the fuel consumption rate in kg/m2s

041202-4 / Vol. 134, APRIL 2012

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/12/2013 Terms of Use: http://asme.org/terms

Fig. 4

Average ame spread velocity versus air velocity

number, taking the pool length and free stream air velocity as the characteristic length and characteristic velocity respectively, falls in the range of 43,000 to 340,000. The ow is laminar in this range, as reported for the same conguration [12]. Time step of 1 105 s is used for the computations. A grid independence study has been conducted as reported in Ali et al. [29]. After setting the initial conditions, the ignition source is added as a localized volumetric heat generation. 4.1 Flame Spread Velocity. Figure 2(a) shows the temporal variation of ame tip location for different air velocities. Clearly, the ame spread is accelerated as the air velocity increases. The initial temperature of the liquid pool is equal to the ambient temperature (300 K), which is higher than the ash point temperature

(284 K) of methanol. Besides, methanol is a high volatile fuel with a low boiling point (338 K). As the air velocity parallel to the surface is increased, there will be an increase in the availability of oxidizer near the pool surface. As a result, the gradient of the fuel mass fraction at the pool surface increases [29]. The evaporation rate is controlled, in general, by the gradient of the mass fraction of the fuel at the surface as per the Ficks law. Further, the ame stand-off is reduced at higher air velocities, since the fuel need not travel a larger distance to mix with required amount of air in those cases. Therefore, due to enhanced heat and mass transfers, the ame will able to heat up the portion of the fuel ahead to a signicant length and facilitate continuous and rapid evaporation. It is also clear from Figure 2(a) that the movement of ame tip along the pool surface is steady, without any oscillations or pulsations, as observed in the cases of opposed ow ame spread [23,25] or when the pool temperature is below the ash point of the fuel [22,25], even though it is not linear. This is due to the combined effect of concurrent ow as well as the initial temperature of the pool (300 K) being larger than the ash point of methanol (284 K). Figure 2(b) shows the temporal variation of integrated fuel consumption rate for different air velocities. It is clear that the fuel consumption rate increases with air velocity. Figure 3 presents the variation of ame spread velocity (Vf) along the pool surface. It is clear that the ame spread velocity is not uniform at all the x-locations. It has a higher value initially due to the effect of ignition and also due to enhanced mass transfer at the leading edge. The ame has an anchoring point close to the leading edge of the fuel pool. As the ame tip propagates, the ame spread velocity is seen to decrease and the trend continues till the ame propagates to around 40% to 50% of the pool length. This is due to an increase in the ame stand-off distance along the pool length. After this, there is a slight increase in the ame spread velocity, when the ame moves further downstream toward

Fig. 5 Temperature contours and velocity vectors as a function of time in the gas-phase for u 5 3.9 m/s

Journal of Heat Transfer

APRIL 2012, Vol. 134 / 041202-5

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/12/2013 Terms of Use: http://asme.org/terms

Fig. 6 Temperature contours and velocity vectors as a function of time in the liquid-phase for u 5 3.9 m/s

almost the trailing edge of the fuel pool. This is due to formation of a secondary ame anchoring point at a particular pool location and its propagation, in addition to the leading edge anchoring point. This trend is explained later along with temperature contours for a particular case. At higher air velocities, there is a slight decrease in the ame spread velocity when the ame approaches the trailing edge of the pool. During this period, the evaporation rate of fuel vapor from the short pool surface ahead of the ame tip is not enough to facilitate ame propagation at the same rate. Figure 4 shows the average ame spread velocity obtained by integrating the curves in Fig. 3. Also shown in Fig. 4 are the corresponding experimental values reported in literature [12]. It is seen that the numerical model is able to predict the average ame spread velocity as a function of air velocity reasonably well. 4.2 Flow and Thermal Fields. Instantaneous gas-phase velocity vector and grayscale temperature plots are shown in Fig. 5 for free stream air velocity of 3.9 m/s. The grayscale temperature contours qualitatively indicate the ame zone with the darkest region indicating the maximum temperature. Velocity vectors are shown in ve sections along the pool length at x/L locations starting with 0.05 and ending with 0.95. At a time instant of 0.05 s, ame has propagated nearly up to 20% of the pool length. The ame is seen to stand closer to the surface. This means signicant heat and mass transfer occurs between the gas and liquid-phases. Due to this, the ame spread velocity is seen to be much higher initially, as shown in Fig. 3. The velocity vector at the x/L location of 0.05 shows velocity overshoot in its prole. At this time instant, u-velocity in the gas-phase does not reach the free stream value at the downstream locations. The corresponding liquid-phase quantities are shown in Fig. 6. Since the velocity and temperature scales are much different in liquid- and gas-phases, these quantities are presented separately using 041202-6 / Vol. 134, APRIL 2012

different length scales for velocity vector dimensions as indicated in Figs. 5 and 6. Figure 6(a) shows the instantaneous ow and thermal elds in the liquid-phase at the time instant of 0.05 s. The pool surface has been heated up to approximately 20% of its total length, the location up to which the ame in the gas-phase has propagated. Due to Marangoni convection and shear stress in the interface, an x-direction velocity has been induced on the pool surface. This is clearly shown by the velocity vectors at x/L location of 0.25. However, due to higher viscosity in the liquid-phase, the induced u-velocity gets damped out quickly along the depth as well as the pool length. Furthermore, at the locations where the ame has crossed, the liquid-phase u-velocity value decreases. This is due to reduction in the shear stress at the interface as well as the reduction in the temperature gradient along the pool surface. At a time instant of 0.15 s (Fig. 5(b)), the ame has propagated nearly half of the pool length. The ame extent in the y-direction has also increased, which affects the heat transfer to the pool surface. During this time-period, the ame spread velocity shows a decreasing trend as shown in Fig. 3. In the x-locations where the ame has propagated (x/L 0.05 and 0.25), the velocity overshoot in the vector prole is clearly observed. At x/L location around 0.5, where the ame tip nearly touches to the pool surface, transverse component of velocity (v-velocity) is induced in the ow eld. Figure 6(b) shows the corresponding liquid-phase thermal and ow elds for this time instant. The surface temperature of the pool has increased up to the point where the ame has propagated, as seen before. Just after the point where the ame tip nearly touches the pool surface, the liquid-phase u-velocity has increased signicantly. At the locations where the ame has crossed, the magnitude of u-velocity is seen to decrease due to the reasons mentioned previously. At a time instant of 0.25 s (Fig. 5(c)), the ame tip has propagated nearly 80% of the pool length. The ame extent in the y-direction has increased, especially to a larger extent, near its tip. Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/12/2013 Terms of Use: http://asme.org/terms

A continuous single ame zone as observed in the previous time instants is not seen in this case. The tip of the ame anchored around the leading edge is seen to shift upward at the downstream locations, due to buoyancy induced ow. A second anchoring location for the ame, closer to the pool surface at x/L around 0.7, is now observed. A secondary ame is seen to propagate. This is the reason for the slight increase in the ame spread velocity as seen in Fig. 3. The ame tip has propagated ahead of this secondary ame anchoring point. As seen in previous cases, at the x-locations where the ame has propagated, the velocity overshoot is observed. At an x/L location around 0.75, transverse component of velocity (v-velocity) is induced in the ow eld as before. Figure 6(c) shows the corresponding quantities in the liquid-phase at this time instant. The surface temperature of the pool has increased up to the point where the ame has its second anchoring point. Just after this point, an increase in the liquidphase u-velocity is also noticed. At a time instant of 0.35 s (Fig. 5(d)), the ame tip has crossed the entire pool length. The downstream anchoring point has moved up to 90% of the pool length. The velocity proles upstream of this location have become almost similar. The ame tip height has increased further due to buoyancy induced ow. At

x/L location around 0.95, transverse component of velocity (v-velocity) is induced and a velocity overshoot is also observed. Figure 6(d) shows the corresponding quantities in the liquid-phase at this time instant. The surface temperature of the pool has increased up to 90% of the pool length. At time instant of 0.5 s (Fig. 5(e)), the ame has propagated across the entire pool length. The downstream anchoring point has also moved away beyond the trailing edge of the pool. All the velocity proles have become almost similar. The ame tip height has reduced as the forced convection effects take over the buoyancy induced ow effects. Figure 6(e) shows the corresponding quantities in the liquid-phase, where the surface temperature has increased over the entire pool length. Signicant decrease in the liquid-phase velocity values is also noticed. The velocity vectors observed at this time instant are due to temperature gradient along the pool surface; the surface temperature decreases from the leading edge of the pool, where the ame anchors much closer to the surface, to its trailing edge, where the ame stand-off is at a higher y-location. 4.3 Temporal Variations of Interface Quantities. Figures 7(a) and 7(b) show the variations of surface temperature as a function of time for two free stream velocity cases. As mentioned

Fig. 7 Variation of interface quantities (a, b) temperature, (c, d) gas-phase u-velocity (x-component) and (e, f) gas-phase v-velocity (y-component) along the pool length for various free stream velocity cases; (a), (c) and (e) for u 5 1.3 m/s; (b), (d), and (f) for u 5 3.9 m/s

Journal of Heat Transfer

APRIL 2012, Vol. 134 / 041202-7

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/12/2013 Terms of Use: http://asme.org/terms

earlier, the initial temperature of the liquid pool is at the ambient temperature of 300 K. In general, since methanol has a low boiling point (338 K), the surface temperature reaches an equilibrium value in the range of approximately 325 K to 328 K, at various time instants. As the ame propagates, it heats up the pool surface to the equilibrium value, up to the point of propagation. After this point, there is a sharp reduction in the surface temperature to the ambient value. This has been discussed with respect to Figs. 5 and 6. At an air velocity as low as 1.3 m/s (Fig. 7(a)), the ame is able to anchor very close to the leading edge of the fuel pool, that is, at x/L 0. Therefore, starting from the leading edge itself, the pool surface has reached the equilibrium value (Fig. 7(a)). However, as the air velocity is increased, for example, to 3.9 m/s (Fig. 7(b)), the initial ame anchoring point near to the leading edge of the fuel pool moves slightly downstream. Due to this, the surface temperature at the leading edge has a lower value (Fig. 7(b)) than the lower velocity case (Fig. 7(a)). It can also be seen that the value of the surface temperature at the leading edge decreases as the air velocity is increased from 1.3 m/s to 3.9 m/s. Further reduction was noticed when the air velocity was increased to 5.1 m/s (not shown). The temperature then rises to its equilibrium value at the location on the pool surface, where the ame has its initial anchoring point for that air velocity. Figures 7(c) and 7(d) present the temporal variations of u-velocity for two air velocity cases. Based on the ame anchoring position, a negative u-velocity is induced near to the leading edge of the fuel pool. At the lower air velocity of 1.3 m/s, when the ame anchors just above the leading edge of the pool, the uvelocity is seen to be mostly positive. It sharply increases to a maximum value in the range of around 0.6 m/s to 0.8 m/s at a location in the pool surface just ahead of the point that is heated up by the ame. At higher velocities, since the initial ame anchoring point is on the downstream location of the leading edge of the fuel pool, there is a positive temperature gradient in the negative x-direction between the point of ame anchoring and the pool leading edge. As a result, a negative u-velocity is induced at those locations. It can also be noted that at higher air velocities, there is a reduction in the induced u-velocity as compared to the case of 1.3 m/s. This is mainly due to increased uniformity in the ame tip propagation at higher air velocities. Figures 7(e) and 7(f) show the temporal variations of gasphase v-velocity for two air velocity cases. This component represents the Stefan velocity induced due to vaporization of the liquid fuel. Based on the mass fraction of the fuel vapor at the interface, it gradient value is calculated. Using this interface fuel vapor mass fraction and its gradient, this velocity is calculated by Ficks Law. At the lower air velocity of 1.3 m/s, when the fuel vapor mass fraction is around its equilibrium value at the leading edge, the v-velocity has its highest value of around 0.0067 m/s at that location. Then, its value decreases along the pool surface. Its value spikes up to a local maximum at various locations on the pool surface at various time instants, based on the ame propagation velocity (or the air velocity). At higher air velocity of 3.9 m/s, the maximum v-velocity is observed at the initial ame anchoring location, which is just ahead of the pool leading edge.

increased, due to increased availability of oxidizer near the pool surface, and this also enhances the ame spread velocity. When ignited, the ame anchors near the leading edge of the pool and the tip propagates. The rate of propagation of ame tip along the surface is seen to be steady without uctuations, since the initial liquid pool temperature is higher than the ash point temperature. The ame spread velocity is found to be nonuniform along the pool surface. It is seen to be higher initially due to the effect of ignition and also due to enhanced mass transfer at the leading edge caused by ame anchoring. It then decreases up to a point when the ame has propagated to around 40% to 50% of the pool length, due to increased ame stand-off from the pool surface. At this position, a secondary ame anchoring point is observed, which propagates toward the trailing edge of the pool. As a result, there is an increasing trend observed in the ame spread velocity during this period. As the air velocity is increased, the initial ame anchoring point moves downstream of the leading edge of the fuel pool. The variations of interface quantities in both gas and liquidphase sides depend on the initial ame anchoring location and the attainment of thermodynamic equilibrium between the liquid- and gas-phases.

Acknowledgment
Authors thank Department of Science and Technology India for funding the purchase of Xeon workstations.

References
[1] Burgoyne, J. H., Roberts, A. F., and Quinton, P., 1968, The Spread of Flame Across a Liquid Surface. I. The Induction Period. II. Steady-State Conditions. III. A Theoretical Model, Proc. R. Soc. A, 308, pp.5568. [2] Glassman, I., and Hansel, J. G., 1968, Some Thoughts and Experiments on Liquid Fuel Spreading, Steady Burning and Ignitability in Quiescent Atmospheres, Fire Res. Abstr. Rev., 10, pp. 217234. [3] Sirignano, W. A., and Glassman, I., 1970, Flame Spreading Above Liquid Fuels: Surface-Tension Driven Flows, Combust. Sci. Technol., 1(4), pp.307312. [4] Sirignano, W. A., 1972, A Critical Discussion of Theories of Flame Spread Across Solid and Liquid Fuels, Combust. Sci. Technol., 6(12), pp.95105. [5] Ray, S. R., Fernandez-Pello, A. C., and Glassman, I., 1980, A Study of the Heat Transfer Mechanisms in Horizontal Flame Propagation, J. Heat Transfer, 102(2), pp. 357364. [6] Akita, K., 1973, Some Problems of Flame Spread Along a Liquid Surface, Proc. 14th Int. Symp. on Combustion, Combustion Institute, Pittsburgh, PA, pp. 10751083. [7] Akita. K., and Fujiwara, O., 1971, Pulsating Flame Spread Along the Surface of Liquid Fuels, Combust. Flame, 17(2), pp.268269. [8] Glassman, I., Dryer, F. L., 1980, Flame Spreading Across Liquid Fuels, Fire Saf. J., 3, pp.123128. [9] Ross, H. D., 1994, Ignition of and Flame Spread Over Laboratory-Scale Pools of Pure Liquid Fuels, Prog. Energy Combust. Sci., 20(1), pp.1763. [10] Wichman, I. S., and Baum, H. R., 1998, An Integral Analysis of Two Simple Model Problems of Wind-Aided Flame Spread, Trans. ASME J. Heat Transfer, 110(2), pp. 437442. [11] Hirano, T., Suzuki, T., Mashiko, I., and Tanabe, N., 1980, Gas Movements in Front of Flames Propagating Across Methanol, Combust. Sci. Technol., 22(12), pp.8391. [12] Suzuki, T., and Hirano, T., 1982, Flame Propagation Across a Liquid Ffuel in an Air Stream, Proc. 19th Int. Symp. on Combustion, Combustion Institute, Pittsburgh, PA, pp. 877884. [13] Takeno, K., and Hirano, T., 1986, Flame Spread Over Porous Solids Soaked With a Combustible Liquid, Proc. 21st Int. Symp. on Combustion, Combustion Institute, Pittsburgh, PA, pp. 7581. [14] Takeno, K., and Hirano, T., 1989, Behavior of Combustible Liquid Soaked in Porous Beds During Flame Spread, Proc. 22nd Int. Symp. on Combustion, Combustion Institute, Pittsburgh, PA, pp. 12231230. [15] Ito, A., Masuda, D., and Saito, K., 1991, A Study of Flame Spread Over Alcohols Using Holographic Interferometry, Combust. Flame, 83(34), pp. 375389. [16] Konishi, T., Ito, A., Kudou, Y., and Saito, K., 2002, The Role of a FlameInduced Liquid Surface Wave on Pulsating Flame Spread, Proc. Combust. Inst. 29, pp.267272. [17] Ito, A., Narumi, A., Konishi, T., Tashtoush, G., Saito, K., and Cremers, C. J., 1999, The Measurement of Transient Two-Dimensional Proles of Velocity and Fuel Concentration Over Liquids, Trans. ASME J. Heat Transfer, 121(2), pp.413420. [18] Ross, H. D., and Miller, F. J., 1998, Flame Spread Across Liquid Pools With Very Low-Speed Opposed or Concurrent Airow, Proc. 27th Int. Symp. on Combustion, Combustion Institute, Pittsburgh, PA, pp. 27232729.

Conclusions

Concurrent ame spread over methanol pool surface under atmospheric conditions and normal gravity has been numerically investigated using a transient, two-phase, reacting ow numerical model. Variable thermo-physical properties, single-step global reaction mechanism with partial equilibrium based dissociation of carbon-dioxide, optically thin approximation based radiation model and thermodynamic equilibrium at the interface forms the salient features of the numerical model. The average ame spread velocities for different concurrent air velocities predicted by the model are quite close to the reported experimental measurement [12]. The fuel consumption rate increases as the air velocity is 041202-8 / Vol. 134, APRIL 2012

Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/12/2013 Terms of Use: http://asme.org/terms

[19] Takahashi, K., Kodaira, Y., Kudo, Y., Ito, A., and Saito, K., 2007, Effect of Oxygen on Flame Spread Over Liquids, Proc. Combust. Inst. 31, pp. 26252631. [20] Torrance, K. E., 1971, Subsurface Flows Preceding Flame Spread Over a Liquid Fuel, Combust. Sci. Technol., 3(3), pp.133143. [21] Torrance, K. E. and Mahajan, R. L., 1975, Surface Tension Flows Induced by a Moving Thermal Source, Combust. Sci. Technol., 10(34), pp. 125136. [22] Epstein, M., and Burelbach, J. P., 1998, Effective Thermal Conductivity Model of Flame Spread Over a Shallow Subash Liquid Fuel Layer, Trans. ASME J. Heat Transfer, 120(3), pp.781785. [23] Di Blasi, C., Crescitelli, S., Russo, G., Cinque, G., 1991, Model of Pulsating Flame Spread Across Liquid Fuels, Proc. 23rd Int. Symp. on Combustion (Pittsburgh, PA: Combustion Institute), pp.16691675. [24] Di Blasi, C., Crescitelli, S., and Russo, G., 1991, Model of Oscillatory Phenomena of Flame Spread Along the Surface of Liquid Fuels, Comput. Methods Appl. Mech. Eng., 90(13), pp. 643657.

[25] Schiller, D. N., Ross, H. D., and Sirignano, W. A., 1996, Computational Analysis of Flame Spread Across Alcohol Pools, Combust. Sci. Technol., 118(4-6), pp. 203255. [26] Schiller, D. N., and Sirignano, W. A., 1997, Opposed-Flow Flame Spread Across n-Propanol Pools, Proc. 26th Int. Symp. on Combustion (Pittsburgh, PA: Combustion Institute), pp. 13191325. [27] Kim, I., and Sirignano, W. A., 2003, Computational Study of Opposed-ForceFlow Flame Spread Across Propanol Pools, Combust. Flame, 132(4), pp. 611627. [28] Kim, I., Schiller, D. N., and Sirignano, W. A., 1998, Axisymmetric Flame Spread Across Propanol Pools in Normal and Zzero Gravities, Combust. Sci. Technol., 139(1), pp. 249275. [29] Ali, S. M., Raghavan, V., and Tiwari, S., 2010, A Study of Steady Laminar Diffusion Flame Over Methanol Pool Surface, Int. J. Heat Mass Transfer, 53(2122), pp. 46964706. [30] Patankar, S. V., Numerical Heat Transfer and Fluid Flow, McGraw-Hill, New York, 1980.

Journal of Heat Transfer

APRIL 2012, Vol. 134 / 041202-9

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 04/12/2013 Terms of Use: http://asme.org/terms

Das könnte Ihnen auch gefallen