Está en la página 1de 9

Diffusion flames and droplet combustion Dr M.

Lawes

Diffusion flames, droplet evaporation and droplet combustion

Introduction
Conventional spark ignition engines operate with premixed combustion in which the fuel has fully
vaporised and mixed with air before flame initiation. Similarly advanced concepts such as Gasoline
direct injection (GDI) and homogeneous charge compression ignition (HCCI) utilise premixed
reactants.
However, it is possible that further advances in engine design, moving to greater stratification, might
result in some droplet combustion, as is the case in diesel and gas turbine engines. Therefore, and to
provide a complement to the laminar and turbulent premixed combustion discussed elsewhere,
diffusion and droplet flames are introduced here as well as droplet evaporation.

Diffusion flames
Before considering liquid fuels, the simpler case of a gas burning in oxidant (usually air) will be
examined e.g. an industrial or power station gas-fired furnace. Consider the elementary case of
gaseous fuel and oxidant supplied in concentric tubes. Before combustion can occur, mixing must
take place to form a combustible mixture. Mixing may occur in two ways,

i) MOLECULAR DIFFUSION. A relatively slow process. Such flames are used when long slow
flames are required for distribution of heat over large areas of a furnace.

ii) TURBULENT MASS TRANSFER (EDDY DIFFUSION). Mass transfer is more rapid than
with molecular diffusion, mixing by ‘lumps’ carried into regions of lower concentration. Higher
thermal output per unit volume results.

Consider first the laminar case, dominated by molecular diffusion as shown in Fig. 1. This has been
considered by several workers (Burke and Schuman, Ind. Engineering. Chem, Vol. 20 p. 998 (1928),
and Hawthorne et al. Third Symposium on Combustion, The Combustion Institute, p 254).

Locus of
flame

Air
Fuel

Figure 1 Diffusion flame Figure 2. Species variation through a diffusion flame at


a fixed height above the fuel jet tube.

Assumptions used in the analysis include species mass conservation and infinitely fast burning at very
restricted fuel/air ratios. The result is a differential equation of motion for the flame, which can be
solved with Bessel functions to give the flame geometry.

D:\My Documents\LECTURES\spark course lawes 2005\lawes atomisation\droplet combustion.DOC 1


Last printed 01/05/2007 13:17:00
Diffusion flames and droplet combustion Dr M. Lawes

Its detailed analysis is beyond the scope of the present lecture course and only graphical results are
presented. Figure 2 shows the concentration distribution of fuel, oxygen and nitrogen for the special
case of equal fuel and air velocities in the vertical direction. For flames with excess air, a shape of the
type shown in Fig. 3(a) is indicated by the analysis and the flame is said to be overventilated. For
excess fuel, a shape of the type shown in Fig. 3(b) results and the flame is underventilated.

Figure 3. Shape of diffusion flames at under and over ventilated conditions.

For a given pair of gases, the height of the flame is found to be proportional to the velocity of flow and
to the square of the diameter, d, of the jets.

In practice, laminar conditions only apply over a limited range. A breakdown to turbulent conditions
occurs at higher gas flow rates. In such flames eddy diffusion is more significant than molecular
diffusion and convective effects become of importance. The flame becomes thicker, resulting in a
greater reaction rate. The basic analysis of such flames follows a similar pattern, albeit more complex,
to that for laminar flames. The height and condition of a diffusion flame as a function of nozzle
velocity is shown in Fig. 4.

Figure 4. Variation of flame height and character as a function of jet velocity

D:\My Documents\LECTURES\spark course lawes 2005\lawes atomisation\droplet combustion.DOC 2


Last printed 01/05/2007 13:17:00
Diffusion flames and droplet combustion Dr M. Lawes

Evaporation of a fuel droplet

Stationary drop
Here we consider the evaporation of a single droplet in a non-convective atmosphere of a given
temperature and pressure as shown in Fig. 5. Following the approach of Spallding, a Quasi-Steady
assumption is used which implies that at a given instant, the droplet radius is considered constant and
the conservation equations do not contain any derivatives w.r.t. time, (hence Quasi-Steady)
Figure 5 evaporating droplet
One may set down, and solve the conservation equations for mass, energy, and the equation of motion.
T
T∞

rl r

Mass rate of evaporation


Heat from atmosphere

The form of solution is dependant on the exact method of solution, but a typical solution, for a Lewis
number (ratio of heat to mass diffusion) of unity, for the mass rate of fuel consumption, m& f , is

4π rl k
m& f = ln(1 + B ) (1)
cp

where c p and k are average specific heat and thermal conductivity at the droplet surface; rl is the
droplet radius and the Transfer Number, B, is given by

B=
1
L
{
c p (T∞ − Tl ) } (2)

where L is the latent heat of evaporation of fuel. Here, the transfer number represents the heat
available due to the difference between droplet and free stream temperatures divided by the heat
required to evaporate the fuel.
Equation (1) gives the mass rate of droplet evaporation for a given droplet diameter. It is more useful
to obtain the droplet diameter as a function of time and the droplet lifetime (time to fully evaporate the
droplet. For this, we consider the conservation of mass. The rate of fuel consumption must be equal to
the time derivative of the total droplet mass:
d ⎛4 3 ⎞
m& f = − ⎜ π rl ρ fl ⎟
dt ⎝ 3 ⎠
4 dr 3 dr
= − π ρf l l l (3)
3 drl dt
drl
= − 4πρ f lrl2
dt

D:\My Documents\LECTURES\spark course lawes 2005\lawes atomisation\droplet combustion.DOC 3


Last printed 01/05/2007 13:17:00
Diffusion flames and droplet combustion Dr M. Lawes

now
d rl2( ) = d (r ) dr
l
2
l
= 2rl
drl
(4)
dt drl dt dt
Substituting Eq. (4) into (1) gives

( )
d rl2
=
− m& f
(5)
dt 2πρ fl rl
Substituting Eq. (1) into (5), and expressing in terms of diameter, gives

( ) = −8k ln(1 + B) = − K
d dl2
(6)
dt cρ fl

where K is a constant. Therefore, during evaporation, the square of the droplet diameter is a linear
function of time; the so-called d2 burning law. K is termed the evaporation constant.
At any instant in time, t, the droplet diameter, d, is given by

∫d d (d l ) = ∫0 − Kdt
d2 2 t
2
0

d 2 = d 02 − Kt (7)

where d 0 is the initial droplet diameter. The droplet lifetime, τ d is given by equating d to 0.

d 02
τd = (8)
K

Drop in a convecting environment


In practice, droplets are usually not stationary, as assumed above. However, it has been
experimentally confirmed that a d2 type law is still applicable. In such a situation, an empirically
derived relation for K is given by

(
K = K 0 10
. + 0.276 Re 0.5 Sc 0.333 ) (9)

−8 k
K0 = ln(1 + B) as in Eq. (6)
c ρ fl

c p (T∞ − Tl )
and B=
L
where K 0 is the evaporation coefficient in a stagnant atmosphere, k is the gaseous thermal
conductivity, ρ l is the density of liquid fuel at Tl , c p is the gaseous specific heat, T is the
ambient gas temperature, Tl is the saturation temperature of the liquid fuel at the local pressure, L is
the sensible enthalpy of liquid fuel from the initial temperature to Tl , plus the latent heat of
evaporation. Re and Sc are the Reynolds and Schmit ( ν/mass diffusion coefficient) numbers. Other
such expressions are available in the literature.

Combustion of a stationary drop of fuel


Knowledge about the combustion of a single fuel droplet forms the building blocks on which more
practical studies can be based. It allows information to be gained on such parameters as evaporation,

D:\My Documents\LECTURES\spark course lawes 2005\lawes atomisation\droplet combustion.DOC 4


Last printed 01/05/2007 13:17:00
Diffusion flames and droplet combustion Dr M. Lawes

mixing and burning in a well characterised environment. Analysis of the combustion of a stationary
drop follows the same procedure as that for an evaporating one. However, the heat generated by
combustion also must be considered. Assumptions are:
1. At a given instant, the droplet radius is considered constant and the conservation equations do not
contain any derivatives w.r.t. time, (hence Quasi-Steady).
2. The flame zone is an infinitesimally thin spherical shell concentric with that of the droplet. This
implies an infinite rate of chemical reaction.
3. Vaporised fuel diffuses from the droplet to the flame zone and oxygen diffuses in towards the
flame zone and neither of them penetrates the flame. The fuel and oxidant react in stoichiometric
proportions at the flame front to form products which diffuse away from the flame zone in both
directions.
4. Heat is transferred to the droplet surface in the amount just sufficient to provide the latent heat of
evaporation. Radiant heat transfer is considered negligible.
With the above assumptions, the mass fractions and temperature profiles are shown in Fig. 6.

Figure 6 Burning of a single stationary droplet


The solution, for a Lewis number (ratio of heat to mass diffusion) of unity is similar to that for an
evaporating droplet with the addition of a combustion term in the transfer number

1⎧ ∆h f O ∞ ⎫
B= ⎨c p (T∞ − Tl ) + ⎬ (10)
L⎩ j ⎭

where L is the latent heat of evaporation of fuel; ∆h the heat of combustion and f O ∞ is the fractional
mass concentration of oxygen away from the droplet. The d2 law still is valid, but the evaporation, or
combustion, constant will be higher than before.

D:\My Documents\LECTURES\spark course lawes 2005\lawes atomisation\droplet combustion.DOC 5


Last printed 01/05/2007 13:17:00
Diffusion flames and droplet combustion Dr M. Lawes

The d2 burning law only holds for a limited range of droplet diameters. For small droplets,
vaporisation is rapid and the flame is almost ‘premixed’ in character. For very large drops, ‘pool
burning’ and the radiation effects ignored in the Q-S theory are significant. The various regimes are
illustrated in Fig. 7.

Characteristic dimension, mm
Figure 7 Regimes of single droplet combustion
The boundaries of these regimes will vary with pressure; e.g. at supercritical pressures, the droplet
will, as soon as the temperature everywhere exceeds the critical temperature, become a puff of gas that
will then burn as a gaseous diffusion flame.

Arrays of Droplets
In reality one is unlikely to have a situation of evaporation or combustion of droplets of a uniform
size. Usually the fuel injection process (e.g. atomiser) will give rise to a spectrum of different droplet
sizes. The resultant droplet size distribution rarely fits any of the usual statistical functions (e.g.
Gaussion). Therefore, it is usual to describe the droplet array in terms of its Sauter Mean Diameter
(SMD). This is defined as the diameter of a droplet whose volume to surface ratio would be the same
as that of the entire spray.

SMD=
∑ nd 3 (11)
∑ nd 2
where n is the number of droplets of diameter d. The SMD, being proportional to the spray volume (or
mass) to surface ratio is an indicator of the degree of atomisation of a particular fuel injector. It is
common to express the performance of an injector in terms of the SMD. For example it is found that
the performance of an aero engine ‘Simplex’ type of atomiser can be correlated over a wide range of
conditions by the relation:
120ν 0.215σ 0.33 ρ 0.2 Q 0.209
SMD = (12)
∆P 0.458
where ν is the kinematic viscosity (centistokes), ρ is the fuel density (gm/cc), Q is the fuel flow
(US gal/hr), ∆P is the injector pressure drop (psi), and σ is the surface tension (kg s-2).

A number of experimental techniques, including high speed photography, have been used to determine
the SMD of a particular spray.

D:\My Documents\LECTURES\spark course lawes 2005\lawes atomisation\droplet combustion.DOC 6


Last printed 01/05/2007 13:17:00
Diffusion flames and droplet combustion Dr M. Lawes

Regimes of droplet combustion


As with single droplet combustion in a quiescent atmosphere, in many situations involving droplet
combustion, mixing and burning can, to a first approximation, be assumed to be very fast (Spalding,
1952). Hence, evaporation is the rate determining process. This is particularly so for large droplets,
and most of the early experiments were conducted on droplets with diameters of the order of
millimetres. However in these cases the d2 burning law may not be a good approximation and other
factors must be considered (Annamalai & Ryan, 1992).

Evaporation based description of cluster burning


The d2 law fails when several drops are close enough to interfere with their individual evaporation and
combustion processes. Research into drop array, cloud or cluster burning has been reviewed by
Annamalai and Ryan (1992). A cloud or cluster is defined as having a large number of droplets,
usually much higher than 10. Droplets in clouds can have a very different behaviour from isolated
ones due to interaction between the droplets, and this depends on the mixture composition within the
cloud. For combustion to take place, liquid fuel must first be vaporised and this results in a complex
mixture of fuel droplets, fuel vapour and air. The overall mixture of fuel and air may or may not be
within its flammability limits since it is the local proportions of vapour and air that determine both
whether the overall mixture is flammable, and the location of the reaction zone.

Conditions in the droplet cloud can be expressed in terms of the liquid volume of fuel present per unit
volume of mixture. This is a function of the average diameter of the droplets, d, and the average
interdrop distance, l . The ratio, d/ l , was first used to express the limiting conditions for flame
propagation, in terms of flammability of the aerosol. This parameter has been developed further, to
accommodate more complex processes in the cloud. Various definitions have been used to describe
the phenomena of evaporation and subsequent combustion conditions in aerosol clouds (Labowski et
al., 1978, Chiu et al., 1977 & 198?, Correa et al., 1982). The expression developed by Chiu et al., to
characterise different group combustion regimes in aerosol clouds, is given by

rate of heat exchange between two phases


G=
rate of energy transport by convection

G = 1.5 Le (1 + 0. 276 Re0.5 Sc0.33) NT0.66 (d/ l ) (13)

where Le is the Lewis number, Sc the Schmidt number, Re is the droplet Reynolds number and NT
denotes the total number of droplets in a cluster. The group combustion number, G, corresponds to the
ratio of the gross droplet evaporation to the inward oxygen diffusion (Akamatsu et al., 1996). As
shown in Fig. 8, four group combustion modes of a droplet cloud are possible as follows:

a) External sheath burning, G > 102. This consists of an inner non-vaporising droplet cloud
surrounded by a vaporising droplet layer with the flame at a "stand-off" distance from the spray
boundary. High-G sprays usually have large group-burning rates and low core temperatures.

b) External group combustion, G > 1. The spray zone consists of an inner vaporising cloud with a
stand-off diffusion flame from the boundary of the droplets.

c) Internal group combustion, 10-2 < G < 1. In this mode, the main flame locates within the spray
boundary, while individual drop burning occurs in the outer regions of the spray.

d) Individual droplet combustion, G < 10-2.

D:\My Documents\LECTURES\spark course lawes 2005\lawes atomisation\droplet combustion.DOC 7


Last printed 01/05/2007 13:17:00
Diffusion flames and droplet combustion Dr M. Lawes

Figure 8 Four group combustion modes

A schematic of liquid-fuel spray group combustion is shown in Fig. 9. In this figure, Chiu and Croke
(1981) subdivided the spray flame into several zones

1) A potential core,
2) external group-combustion zone with evaporating droplets,
3) turbulent envelope diffusion flame at spray core boundary,
4) multi-droplet combustion zone with internal group-combustion behaviour, and
5) turbulent brush flame.

Chiu and Croke used group-combustion theory to conduct predictive calculations for a C10H14 spray
flame. They predicted temperature and concentration profiles that indicate that the flame is stabilised
near the spray boundary. They also reported that a relative minimum in temperature profile occurs on
the axis of the spray, where the fuel vapour concentration has a maximum.

Figure 9 Schematic of liquid-fuel spray group combustion

D:\My Documents\LECTURES\spark course lawes 2005\lawes atomisation\droplet combustion.DOC 8


Last printed 01/05/2007 13:17:00
Diffusion flames and droplet combustion Dr M. Lawes

Supercritical evaporation/burning of liquid droplets


At high pressures, the above analyses become inaccurate and are nearly useless above the critical
pressure. Shown in Fig. 10 are phase diagrams: P-v and P-T. At low (normal) pressures there exists a
clear boundary between the liquid and vapour phases. However, above the critical conditions, no such
boundary exists. (For n-octane, the critical pressure and temperature are 2.49 MPa and 568.8 K).
There are two main reasons why the above analysis is incorrect at high pressures which might prevail
in some direct injection engines. First, as the pressure approaches the critical pressure, the latent heat
of vaporization, L, approaches zero and Eqs. (1) and (2) yields an evaporation rate that rises to infinity.
Second, the gaseous mass in the region influenced by the droplet increases with pressure so that
derivatives w.r.t. time in the energy equation cannot be neglected.

Figure 10. Pressure v volume and pressure v temperature curves for a liquid fuel

An analysis of supercritical droplet burning is beyond the scope of the present lecture. However, it is
important to understand that when critical conditions are approached or exceeded, very fast
evaporation (flash boiling) will result. Although these conditions don’t occur in current engines,
(because fuel injection occurs at lower pressure before ignition and before top dead centre), they might
exist if, for example, fuel where to be injected after the start of combustion.

D:\My Documents\LECTURES\spark course lawes 2005\lawes atomisation\droplet combustion.DOC 9


Last printed 01/05/2007 13:17:00

También podría gustarte