combustion || diffusion flames

61
Diffusion flames 6 6.1 INTRODUCTION Earlier chapters were concerned with flames in which the fuel and oxidizer are ho- mogeneously mixed. Even if the fuel and oxidizer are separate entities in the initial stages of a combustion event and mixing occurs rapidly compared to the rate of com- bustion reactions, or if mixing occurs well ahead of the flame zone (as in a Bunsen burner), the burning process may be considered in terms of homogeneous premixed conditions. There are also systems in which the mixing rate is slow relative to the reaction rate of the fuel and oxidizer, in which case the mixing controls the burning rate. Most practical systems are mixing-rate-controlled and lead to diffusion flames in which fuel and oxidizer come together in a reaction zone through molecular and turbulent diffusion. The fuel may be in the form of a gaseous fuel jet or a condensed medium (either liquid or solid), and the oxidizer may be a flowing gas stream or the quiescent atmosphere. The distinctive characteristic of a diffusion flame is that the burning (or fuel consumption) rate is determined by the rate at which the fuel and oxidizer are brought together in proper proportions for reaction. Since diffusion rates vary with pressure and the rate of overall combustion reac- tions varies approximately with the pressure squared, at very low pressures the flame formed will exhibit premixed combustion characteristics even though the fuel and oxidizer may be separate concentric gaseous streams. Figure 6.1 details how the flame structure varies with pressure for such a configuration where the fuel is a simple higher-order hydrocarbon [1]. Normally, the concentric fuel-oxidizer configuration is typical of diffusion flame processes. 6.2 GASEOUS FUEL JETS Diffusion flames have far greater practical application than premixed flames. Gaseous diffusion flames, unlike premixed flames, have no fundamental character- istic property, such as flame velocity, which can be measured readily; even initial mixture strength (the overall oxidizer-to-fuel ratio) has no practical meaning. Indeed, a mixture strength does not exist for a gaseous fuel jet issuing into a quies- cent atmosphere. Certainly no mixture strength exists for a single small fuel droplet burning in the infinite reservoir of a quiescent atmosphere. CHAPTER Combustion. http://dx.doi.org/10.1016/B978-0-12-407913-7.00006-2 Copyright © 2015 Elsevier Inc. All rights reserved. 301

Upload: irvin

Post on 11-Apr-2017

231 views

Category:

Documents


9 download

TRANSCRIPT

Page 1: Combustion || Diffusion flames

Diffusion flames 66.1 INTRODUCTIONEarlier chapters were concerned with flames in which the fuel and oxidizer are ho-mogeneously mixed. Even if the fuel and oxidizer are separate entities in the initialstages of a combustion event and mixing occurs rapidly compared to the rate of com-bustion reactions, or if mixing occurs well ahead of the flame zone (as in a Bunsenburner), the burning process may be considered in terms of homogeneous premixedconditions. There are also systems in which the mixing rate is slow relative to thereaction rate of the fuel and oxidizer, in which case the mixing controls the burningrate. Most practical systems are mixing-rate-controlled and lead to diffusion flamesin which fuel and oxidizer come together in a reaction zone through molecular andturbulent diffusion. The fuel may be in the form of a gaseous fuel jet or a condensedmedium (either liquid or solid), and the oxidizer may be a flowing gas stream orthe quiescent atmosphere. The distinctive characteristic of a diffusion flame isthat the burning (or fuel consumption) rate is determined by the rate at which thefuel and oxidizer are brought together in proper proportions for reaction.

Since diffusion rates vary with pressure and the rate of overall combustion reac-tions varies approximately with the pressure squared, at very low pressures the flameformed will exhibit premixed combustion characteristics even though the fuel andoxidizer may be separate concentric gaseous streams. Figure 6.1 details howthe flame structure varies with pressure for such a configuration where the fuel isa simple higher-order hydrocarbon [1]. Normally, the concentric fuel-oxidizerconfiguration is typical of diffusion flame processes.

6.2 GASEOUS FUEL JETSDiffusion flames have far greater practical application than premixed flames.Gaseous diffusion flames, unlike premixed flames, have no fundamental character-istic property, such as flame velocity, which can be measured readily; even initialmixture strength (the overall oxidizer-to-fuel ratio) has no practical meaning.Indeed, a mixture strength does not exist for a gaseous fuel jet issuing into a quies-cent atmosphere. Certainly no mixture strength exists for a single small fuel dropletburning in the infinite reservoir of a quiescent atmosphere.

CHAPTER

Combustion. http://dx.doi.org/10.1016/B978-0-12-407913-7.00006-2

Copyright © 2015 Elsevier Inc. All rights reserved.301

Page 2: Combustion || Diffusion flames

6.2.1 APPEARANCEOnly the shape of the burning laminar fuel jet depends on the mixture strength. If ina concentric configuration the volumetric flow rate of air flowing in the outerannulus is in excess of the stoichiometric amount required for the volumetricflow rate of the inner fuel jet, the flame that develops takes a closed, elongatedform. A similar flame forms when a fuel jet issues into the quiescent atmosphere.Such flames are referred to as being overventilated. If in the concentric configura-tion the air supply in the outer annulus is reduced below an initial mixture strengthcorresponding to the stoichiometric required amount, a fan-shaped, underventi-lated flame is produced. The general shapes of the underventilated and overventi-lated flame are shown in Figure 6.2 and are generally referred to as co-flowconfigurations. As will be shown later in this chapter, the actual heights varywith the flow conditions.

The axial symmetry of the concentric configuration shown in Figure 6.2 is notconducive to experimental analyses, particularly when some optical diagnostic toolsor thermocouples are used. There are parametric variations in the r and y coordinatesshown in Figure 6.2. To facilitate experimental measurements on diffusion flames,the so-called WolfhardeParker two-dimensional gaseous fuel jet burner is used.Such a configuration is shown in Figure 6.3, taken from Smyth et al. [2]; the screensshown in the figure are used to stabilize the flame. As can be seen in Figure 6.3,ideally there are no parametric variations along the length of the slot.

1cm

60 mm

30 mm

140 mm

Remainder ofreaction zone

Reaction zone

Luminous zonewith intenselyluminous zone

37 mm

Reaction zone

Radiation fromcarbon particles(luminous zone)

First weakcontinuousluminosityReaction zone

(C2 and CHbands)

Luminous zone

FIGURE 6.1

Structure of an acetyleneeair diffusion flame at various pressures in mmHg.

After Gaydon and Wolfhard [1].

302 CHAPTER 6 Diffusion flames

Page 3: Combustion || Diffusion flames

Other types of gaseous diffusion flames are those in which the flow of the fueland oxidizer oppose each other; these are referred to as counter-flow diffusionflames. The types most frequently used are shown in Figure 6.4. Although these con-figurations are somewhat more complex to establish experimentally, they have def-inite advantages. The opposed jet configuration, in which the reactant streams areinjected as shown in Figure 6.4 as Types I and II, has two major advantages

rj

rs

Air

Fuel

Air

yF Overventilated

Underventilated yF

FIGURE 6.2

Appearance of gaseous fuel jet flames in a co-flow cylindrical configuration.

L1

L2

z

xy

Laser

Stabilizingscreens

Monochromator

P

Flame zones

AirAir

CH4

FIGURE 6.3

Two-dimensional WolfhardeParker fuel jet burner flame configuration.

After Smyth et al. [2].

6.2 Gaseous Fuel Jets 303

Page 4: Combustion || Diffusion flames

compared to co-flow fuel jet or WolfhardeParker burners. First, there is littlepossibility of oxidizer diffusion into the fuel side through the quench zone at thejet tube lip. Second, the flow configuration is more amenable to mathematicalanalysis.

Although the aerodynamic configuration designated as Types III and IV producethe stagnation point counter-flow diffusion flames shown, it is somewhat moreexperimentally challenging. More interestingly, the stability of the process ismore sensitive to the flow conditions. Generally the mass flow rate through theporous media is very much smaller than the free stream flow. Usually the fuel bleedsthrough the porous medium, and the oxidizer is the free stream component.Of course, the reverse could be employed in such an arrangement. However, it isdifficult to distinguish the separation of the flame and flow stagnation planes ofthe opposing flowing streams. Both, of course, are usually close to the porousbody. This condition is alleviated by the approach shown by Types I and II inFigure 6.4.

As one will note from these figures, for a hydrocarbon fuel opposing an oxidizerstream (generally air), the soot formation region that is created and the flame front

FlameFlame

FlameFlame

Matrix

Poroussphere

Porouscylinder

Air Air

Type III Type IV

Type II

FuelType I

Fuel

AirOxidant

Fuel

FIGURE 6.4

Various counter-flow diffusion flame experimental configurations.

304 CHAPTER 6 Diffusion flames

Page 5: Combustion || Diffusion flames

can lie on either side of the flow stagnation plane. In essence, although most of thefuel is diverted by the stagnation plane, some molecules diffuse through the stagna-tion plane to create the flame on the oxidizer side. With the diffusion flame locatedon the oxidizer side, the gas flow velocity (and thermophoresis) directs the soot par-ticles, which are always formed on the fuel side of the flame, away from the high-temperature region of the diffusion flame. Because soot particles approximatelyfollow the gas-phase streamlines, they are not oxidized. For the contra condition,when the diffusion flame and soot region reside on the fuel side of the stagnationplane, the soot particles are carried toward the high-temperature diffusion flameby the gas and thermophoretic velocities and have the potential to oxidize after pas-sage through the flame. In the overventilated co-flow diffusion flame of Figure 6.2,the soot particles also move toward the flame and thus have similar path and temper-ature histories as when the counter-flow diffusion flame is located on the fuel side ofthe stagnation plane. In fact, a liquid hydrocarbon droplet burning in a quiescentenvironment mimics this configuration. With the diffusion flame on the oxidizerside, the soot formation and particle growth are very much different from any realcombustion application.

As Kang et al. [3] have reported, counter-flow diffusion flames are located on theoxidizer side when hydrocarbons are the fuel. Appropriate dilution with inert gasesof both the fuel and oxidizer streams, frequently used in the co-flow situation, canposition the flame on the fuel side. It has been shown [4] that the criterion for theflame to be located on the fuel side is."

mO;0

�Le

1=2O

mF;0

�Le

1=2F

#> i

where mO,0 is the free stream mass fraction of the oxidizer, mF,0 is the free streammass fraction of the fuel, i is the mass stoichiometric coefficient, and Le is the appro-priate Lewis number (see Chapter 4, Section 4.3.2). The Le for oxygen is 1 and forhydrocarbons it is normally greater than 1.

The profiles designating the flame height yF in Figure 6.2 correspond to the stoi-chiometric adiabatic flame temperature as developed in Section 6.2.4. Indeed, thecurved shape designates the stoichiometric adiabatic flame temperature isotherm.However, when one observes an overventilated co-flow flame in which a hydrocar-bon constitutes the fuel jet, one observes that the color of the flame is distinctlydifferent from that of its premixed counterpart. Whereas a premixed hydrocarbon-air flame is violet or blue green, the corresponding diffusion flame varies from brightyellow or orange. The color of the hydrocarbon diffusion flame arises from the for-mation of soot in the fuel of the jet flame. The soot particles then flow through thereaction zone and reach the flame temperature. Some continue to burn through andafter the flame zone and radiate. Due to the temperature that exists and the sensitivityof the eye to various wavelengths in the visible region of the electromagnetic spec-trum, hydrocarbon-air diffusion flames, particularly those of co-flow structure,appear to be yellow or orange.

6.2 Gaseous Fuel Jets 305

Page 6: Combustion || Diffusion flames

As will be discussed extensively in Chapter 8, Section 8.5, the hydrocarbon fuelpyrolyzes within the jet and a small fraction of the fuel forms soot particles that growin size and mass before entering the flame. The soot continues to burn as it passesthrough the stoichiometric isotherm position that designates the fueleair reactionzone and the true flame height. These high-temperature particles continue to burnand stay luminous until they are consumed in the surrounding flowing air. Theend of the luminous yellow or orange image designates the soot burnout distanceand not what one would call the stoichiometric flame temperature isotherm. Forhydrocarbon diffusion flames, the visual distance from the jet port to the end ofthe “orange or yellow” flame is determined by the mixing of the burned gas contain-ing soot with the overventilated air flow. Thus the faster the velocity of the co-flowair stream, the shorter the distance of the particle burnout. Simply, the particleburnout is not controlled by the reaction time of oxygen and soot, but by the mixingtime of the hot exhaust gases and the overventilated gas stream. Thus, hydrocarbonfuel jets burning in quiescent atmospheres appear longer than in an overventilatedcondition. Non-sooting diffusion flames, such as those found with H2, CO, andmethanol, are mildly visible and look very much like their premixed counterparts.Their true flame height can be estimated visually.

6.2.2 STRUCTUREUnlike premixed flames, which have a very narrow reaction zone, diffusionflames have a wider region over which the composition changes and chemicalreaction can take place. Obviously, these changes are due principally to someinterdiffusion of reactants and products. Hottel and Hawthorne [5] were thefirst to make detailed measurements of species distributions in a concentric laminarH2-air diffusion flame. Figure 6.5 shows the type of results they obtainedfor a radial distribution at a height corresponding to a cross-section of the

FIGURE 6.5

Species variations through a gaseous fueleair diffusion flame at a fixed height above a fuel jet

tube.

306 CHAPTER 6 Diffusion flames

Page 7: Combustion || Diffusion flames

overventilated flame depicted in Figure 6.2. Smyth et al. [2] made very detailedand accurate measurements of temperature and species variation across aWolfhardeParker burner in which methane was the fuel. Their results are shownin Figures 6.6 and 6.7.

The flame front can be assumed to exist at the point of maximum temperature,and indeed this point corresponds to that at which the maximum concentrations ofmajor products (CO2 and H2O) exist. The same type of profiles would exist for asimple fuel jet issuing into quiescent air. The maxima arise due to diffusion ofreactants in a direction normal to the flowing streams. It is most important to realizethat, for the concentric configuration, molecular diffusion establishes a bulk veloc-ity component in the normal direction. In the steady state, the flame front produces aflow outward, molecular diffusion establishes a bulk velocity component in thenormal direction, and oxygen plus a little nitrogen flows inward toward the center-line. In the steady state, the total volumetric rate of products is usually greater thanthe sum of the other two. Thus, the bulk velocity that one would observe movesfrom the flame front outward. The oxygen between the outside stream and the flamefront then flows in the direction opposite to the bulk flow. Between the centerlineand the flame front, the bulk velocity must, of course, flow from the centerline out-ward. There is no sink at the centerline. In the steady state, the concentration of theproducts reaches a definite value at the centerline. This value is established by the

Mol

e fr

actio

n

1.0

0.5

0.0�10 �5 0 5 10

N2

O2

CH4

H2O

∑x1

Lateral position (mm)

FIGURE 6.6

Species variations throughout a WolfhardeParker methaneeair diffusion flame.

After Smyth et al. [2].

6.2 Gaseous Fuel Jets 307

Page 8: Combustion || Diffusion flames

diffusion rate of products inward and the amount transported outward by the bulkvelocity.

Since total disappearance of reactants at the flame front would indicate infinitelyfast reaction rates, a more likely graphical representation of the radial distribution ofreactants should be that given by the dashed lines in Figure 6.5. To stress this point,the dashed lines are drawn to grossly exaggerate the thickness of the flame front.Even with finite reaction rates, the flame front is quite thin. The experimental resultsshown in Figures 6.6 and 6.7 indicate that in diffusion flames the fuel and oxidizerdiffuse toward each other at rates that are in stoichiometric proportions. Since theflame front is a sink for both the fuel and oxidizer, intuitively one would expectthis most important observation. Independent of the overall mixture strength, sincefuel and oxidizer diffuse together in stoichiometric proportions, the flame tempera-ture closely approaches the adiabatic stoichiometric flame temperature. It is prob-ably somewhat lower due to finite reaction rates, that is, approximately 90% ofthe adiabatic stoichiometric value [6] whether it is a hydrocarbon fuel or not. Thisobservation establishes an interesting aspect of practical diffusion flames in thatfor an adiabatic situation two fundamental temperatures exist for a fuel: one that cor-responds to its stoichiometric value and occurs at the flame front, and one that occursafter the products mix with the excess air to give an adiabatic temperature that cor-responds to the initial mixture strength.

Mol

e fr

actio

n �

10

1.0

0.5

0.0�10 �5 0 5 10

�2

�1

0

1

2

CO2

CO

H2

Lateral position (mm)

log

(loca

l equ

ival

ence

rat

io)

Φ

FIGURE 6.7

Additional species variations for the conditions of Figure 6.6.

After Smyth et al. [2].

308 CHAPTER 6 Diffusion flames

Page 9: Combustion || Diffusion flames

6.2.3 THEORETICAL CONSIDERATIONSThe theory of premixed flames essentially consists of an analysis of factors such asmass diffusion, heat diffusion, and the reaction mechanisms as they affect the rate ofhomogeneous reactions taking place. Inasmuch as the primary mixing processes offuel and oxidizer appear to dominate the burning processes in diffusion flames, thetheories emphasize the rates of mixing (diffusion) in deriving the characteristics ofsuch flames.

It can be verified easily by experiments that in an ethylene-oxygen premixedflame, the average rate of consumption of reactants is about 4 mol/cm3s, whereasfor the diffusion flame (by measurement of flow, flame height, and thickness ofreaction zoneda crude, but approximately correct approach), the average rate ofconsumption is only 6� 10�5 mol/cm3s. Thus, the consumption and heat releaserates of premixed flames are much larger than those of pure mixing-controlled diffu-sion flames.

The theoretical solution to the diffusion flame problem is best approached inthe overall sense of a steady flowing gaseous system in which both the diffusionand chemical processes play a role. Even in the burning of liquid droplets, a fuelflow due to evaporation exists. This approach is much the same as that presentedin Chapter 4, Section 4.3.2, except that the fuel and oxidizer are diffusing inopposite directions and in stoichiometric proportions relative to each other. Ifone selects a differential element along the x-direction of diffusion, the conserva-tion balances for heat and mass may be obtained for the fluxes, as shown inFigure 6.8.

In Figure 6.8, j is the mass flux as given by a representation of Fick’s law whenthere is bulk movement. From Fick’s law:

j ¼ �DðvrA=vxÞ

A v �d ( Av)

dxΔx

dcpT

dxΔxv (cpT ) v (cpT ) � v

dj

dxΔxj �

dq

dxΔx

Δx

q �

j ≡ �D

q ≡

mA ≡ A /

dmA

dx

�λcp

d(cpT )

dxq

j

mA

H

Av

ρ

ρ

ρ ρ

ρ ρ

ρ ρ

ρ

FIGURE 6.8

Balances across a differential element within a diffusion flame.

6.2 Gaseous Fuel Jets 309

Page 10: Combustion || Diffusion flames

As will be shown in Section 6.3.2, the following form of j is exact (however, thesame form can be derived if it is assumed that the total density does not vary with thedistance x, as of course it actually does):

j ¼ �rDvðrA=rÞ

vx¼ �rD

vmA

vx

where mA is the mass fraction of species A. In Figure 6.8, q is the heat flux given byFourier’s law of heat conduction; _mA is the rate of decrease of mass of species A inthe volumetric element (Dx$1)(g/cm3s), and _H is the rate of chemical energy releasein the volumetric element (Dx$1)(cal/cm3s).

With the preceding definitions, for the one-dimensional problem defined inFigure 6.8, the expression for conservation of a species A (say the oxidizer) is:

vrA

vt¼ v

vx

�ðrDÞ vmA

vx

�� vðrAyÞ

vx� _mA (6.1)

where r is the total mass density, rA is the partial density of species A, and y is thebulk velocity in direction x. Solving this time-dependent diffusion flame problem isoutside the scope of this text. Indeed, most practical combustion problems have asteady fuel mass input. Thus, for the steady problem, which implies steady massconsumption and flow rates, one may not only take vrA/vt as zero but also usethe following substitution:

dðrAyÞdx

¼ d½ðryÞðrA=rÞ�dx

¼ ðryÞ dmA

dx(6.2)

The term (ry) is a constant in the problem since there are no sources or sinks.With the further assumption from simple kinetic theory that Dr is independent oftemperature, and hence of x, Eqn (6.1) becomes:

rDd2mA

dx2� ðryÞ dmA

dx¼ _mA (6.3)

Obviously, the same type of expression must hold for the other diffusing speciesB (say the fuel), even if its gradient is opposite to that of A, so that:

rDd2mB

dx2� ðryÞ dmB

dx¼ _mB ¼ i _mA (6.4)

where _mB is the rate of decrease of species B in the volumetric element (Dx$1) and iis the mass stoichiometric coefficient:

ihmB=mA

The energy equation evolves as it did in Chapter 4, Section 4.3.2, to give:

l

cp

d2�cpT�

dx2� ðryÞ d

�cpT�

dx¼ þ _H ¼ �i _mAH (6.5)

310 CHAPTER 6 Diffusion flames

Page 11: Combustion || Diffusion flames

where _H is the rate of chemical energy release per unit volume, and H is the heatrelease per unit mass of fuel consumed (in joules per gram), that is:

� _mBH ¼ _H; � i _mAH ¼ _H (6.6)

since _mB must be negative for heat release (exothermic reaction) to take place.Although the form of Eqns (6.3)e(6.5) is the same as that obtained in dealing

with premixed flames, an important difference lies in the boundary conditions thatexist. Furthermore, in comparing Eqns (6.3) and (6.4) with Eqns (4.31) and(4.32), one must realize that in Chapter 4, the mass change symbol _u was alwaysdefined as a negative quantity.

Multiplying Eqn (6.3) by iH, and then combining it with Eqn (6.5) for the con-dition Le¼ 1 or rD¼ (l/cp), one obtains:

rDd2

dx2�cpT þ imAH

�� �ry� ddx

�cpT þ imAH

� ¼ 0 (6.7)

This procedure is sometimes referred to as the SchvabeZeldovich transforma-tion. Mathematically, what has been accomplished is that the nonhomogeneousterms ( _m and _H) have been eliminated and a homogeneous differential equation(Eqn (6.7)) has been obtained.

The equations could have been developed for a generalized coordinate system. Ina generalized coordinate system, they would have the form:

V$�ðryÞ�cpT�� �l�cp�V�cpT�� ¼ � _H (6.8)

V$�ðryÞ�mj

�� rDVmj

� ¼ þ _mj (6.9)

These equations could be generalized even further (see Ref. [7]) by simplywriting

Phoj _mj instead of _H, where hoj is the heat of formation per unit mass at

the base temperature of each species j. However, for notation simplicitydandbecause energy release is of most importance for most combustion and propulsionsystemsdan overall rate expression for a reaction of the type which follows willsuffice:

Fþ fO/P (6.10)

where F is the fuel, O the oxidizer, P the product, and f the molar stoichiometricindex. Then Eqns (6.8) and (6.9) may be written as:

V$

"�ry� mj

MWjnj� ðrDÞV mj

MWjnj

#¼ _M (6.11)

V$

"ðryÞ cpT

H MWjnj� ðrDÞV cpT

H MWjnj

#¼ _M (6.12)

6.2 Gaseous Fuel Jets 311

Page 12: Combustion || Diffusion flames

where MW is the molecular weight, _M ¼ _mj=MWjnj; nj ¼ f for the oxidizer, andnj¼ 1 for the fuel. Both equations have the form:

V$��ry�a� �rDVa�� ¼ _M (6.13)

where aT¼ cpT/(H MWjnj) and aj¼mj/(MWjnj). They may be expressed as:

LðaÞ ¼ _M (6.14)

where the linear operation L(a) is defined as:

LðaÞ ¼ V$½ðryÞa� ðrDÞVa� (6.15)

The nonlinear term may be eliminated from all but one of the relationshipsLðaÞ ¼ _M For example:

Lða1Þ ¼ _M (6.16)

can be solved for a1, then the other flow variables can be determined from the linearequations for a simple coupling function U so that:

LðUÞ ¼ 0 (6.17)

where U¼ (aj� al)hUj (js 1). Obviously if 1¼ fuel and there is a fueleoxidizersystem only, j¼ 1 gives U¼ 0 and shows the necessary redundancy.

6.2.4 THE BURKEeSCHUMANN DEVELOPMENTWith the development in the previous section, it is now possible to approach the clas-sical problem of determining the shape and height of a burning gaseous fuel jet in acoaxial stream as first described by Burke and Schumann and presented in detail inLewis and von Elbe [8].

This description is given by the following particular assumptions:

1. At the port position, the velocities of the air and fuel are considered constant,equal, and uniform across their respective tubes. Experimentally, this conditioncan be obtained by varying the radii of the tubes (see Figure 6.2). The molar fuelrate is then given by the radii ratio:

r2j

r2s � r2j

2. The velocity of the fuel and air up the tube in the region of the flame is the sameas the velocity at the port.

3. The coefficient of interdiffusion of the two gas streams is constant.

Burke and Schumann [9] suggested that the effects of assumptions 2 and 3compensate for each other, thereby minimizing errors. Although D increases as

312 CHAPTER 6 Diffusion flames

Page 13: Combustion || Diffusion flames

T1.67 and velocity increases as T, this disparity should not be the main objection.The main objection should be the variation ofDwith T in the horizontal directiondue to heat conduction from the flame.

4. Interdiffusion is entirely radial.5. Mixing is by diffusion only, that is, there are no radial velocity components.6. Of course, the general stoichiometric relation prevails.

With these assumptions, one may readily solve the coaxial jet problem. The onlydifferential equation that one is obliged to consider is:

LðUÞ ¼ 0 with U ¼ aF � aO

where aF¼þmF/MWFnF and aO¼þmO/MWOnO. In cylindrical coordinates, thegeneration equation becomes:

y

D

vU

vy

� 1

r

v

vr

rvU

vr

¼ 0 (6.18)

The cylindrical coordinate terms in v/vq are set equal to zero because of thesymmetry. The boundary conditions become:

U ¼ mF;0

MWFnFat y ¼ 0; 0 � r � rj

¼ � mO;0

MWOnOat y ¼ 0; rj � r � rs

and vU=vr ¼ 0 at r¼ rs, y> 0. In order to achieve some physical insight to thecoupling function U, one should consider it as a concentration or, more exactly, amole fraction. At y¼ 0 the radial concentration difference is:

Ur< rj � Ur> rj ¼mF;0

MWFnF� �mO;0

MWOnO

Ur< rj � Ur> rj ¼mF;0

MWFnFþ mO;0

MWOnO

This difference in U reveals that the oxygen acts as it were a negative fuel con-centration in a given stoichiometric proportion, or vice versa. This result is, ofcourse, a consequence of the choice of the coupling function and the assumptionthe fuel and oxidizer approach each other in stoichiometric proportion.

It is convenient to introduce dimensionless coordinates:

xhr

rs; hh

yD

yr2s

and to define parameters ch rj/rs, an initial molar mixture strength n:

nhmO;0MWFnF

mF;0MWOnO

6.2 Gaseous Fuel Jets 313

Page 14: Combustion || Diffusion flames

and the reduced variable:

g ¼ U

MWFnF

mF;0

Equation (6.18) and the boundary conditions then become:

vg

vh¼1

x

v

vx

xvg

vx

g ¼ 1 at h ¼ 0; 0 � x < c;

g ¼ �n at h ¼ 0; c � x < 1

and:

vg

vx¼ 0 at x ¼ 1; h > 0 (6.19)

Equation (6.19), with these new boundary conditions, has the known solution:

g ¼ ð1þ nÞc2 � nþ 2ð1þ nÞcXNn¼1

1

fn

(J1ðcfnÞ½J0ðfnÞ�2

)

�J0�fnx�exp��f2

nh� (6.20)

where J0 and J1 are Bessel functions of the first kind (of order 0 and 1, respectively),and the fn represent successive roots of the equation J1 (f)¼ 0 (with the orderingconvention fn> fn�1, f0¼ 0). This equation gives the solution for U in the presentproblem.

The flame shape is defined at the point where the fuel and oxidizer disappear, thatis, the point that specifies the place where U¼ 0. Hence, setting g¼ 0 provides arelation between x and h that defines the locus of the flame surface.

The equation for the flame height is obtained by solving Eqn (6.20) for h aftersetting x¼ 0 for the overventilated flame and x¼ 1 for the underventilated flame(also g¼ 0).

The resulting equation is still very complex. Since flame heights are large enoughto cause the factor expð�f2

nhÞ to decrease rapidly as n increases at these values of h,it usually suffices to retain only the first few terms of the sum in the basic equationfor this calculation. Neglecting all terms except n¼ 1, one obtains the roughapproximation:

h ¼

1

f21

!ln

�2ð1þ nÞcJ1ðcf1Þ

½n� ð1þ nÞc�f1J0ðf1Þ�

(6.21)

for the dimensionless height of the overventilated flame. The first zero of J1 (f) isf1¼ 3.83.

The flame shapes and heights predicted by such expressions (see Figure 6.9) areshown by Lewis and von Elbe [8] to be in good agreement with experimentalresultsdsurprisingly so, considering the basic drastic assumptions.

314 CHAPTER 6 Diffusion flames

Page 15: Combustion || Diffusion flames

Indeed, it should be noted that Eqn (6.21) specifies that the dimensionless flameheight can be written as:

h ¼ f ðc; nÞ yFD

yr2s¼ f ðc; nÞ (6.22)

Thus since c¼ (rj/rs), the flame height can also be represented by:

yF ¼ yr2jDc2

f ðc; nÞ ¼ pr2j y

pDf 0ðc; nÞ ¼ Q

pDf 0ðc; nÞ (6.23)

where y is the average fuel velocity, Q is the volumetric flow rate of the fuel, andf0 ¼ (f/c2). Thus, one observes that the flame height of a circular fuel jet is directlyproportional to the volumetric flow rate of the fuel.

In a pioneering paper [10], Roper vastly improved on the BurkeeSchumannapproach to determine flame heights not only for circular ports but also for squareports and slot burners. Roper’s work is significant because he used the fact thatthe BurkeeSchumann approach neglects buoyancy and assumes the mass velocityshould everywhere be constant and parallel to the flame axis to satisfy continuity[7], and then pointed out that resulting errors cancel for the flame height of a circularport burner but not for the other geometries [10,11]. In the BurkeeSchumann anal-ysis, the major assumption is that the velocities are everywhere constant and parallelto the flame axis. Considering the case in which buoyancy forces increase the massvelocity after the fuel leaves the burner port, Roper showed that continuity dictatesdecreasing streamline spacing as the mass velocity increases. Consequently, all vol-ume elements move closer to the flame axis, the widths of the concentration profilesare reduced, and the diffusion rates are increased.

1.4

1.2

1.0

0.8

0.6

0.4

0.2

00 0.2 0.4 0.6 0.8 1.0

0.1

0.2

0.3

0.4

0

0.5

Radial distance, x (in.)

Inch

es o

verv

entil

ated

, y

Inch

es u

nder

vent

ilate

d, y

FIGURE 6.9

Flame shapes as predicted by BurkeeSchumann theory for cylindrical fuel jet systems.

After Burke and Schumann [9].

6.2 Gaseous Fuel Jets 315

Page 16: Combustion || Diffusion flames

Roper [10] also showed that the velocity of the fuel gases is increased due toheating, and that the gases leaving the burner port at temperature T0 rapidly attaina constant value Tf in the flame regions controlling diffusion; thus the diffusivityin the same region is:

D ¼ D0ðTf=T0Þ1:67

where D0 is the ambient or initial value of the diffusivity. Then, considering the ef-fect of temperature on the velocity, Roper developed the following relationship forthe flame height:

yF ¼ Q

4pD0

1

ln½1þ ð1=SÞ�T0Tf

0:67

(6.24)

where S is the stoichiometric volume rate of air to volume rate of fuel. AlthoughRoper’s analysis does not permit calculation of the flame shape, it does producefor the flame height a much simpler expression than Eqn (6.21).

If due to buoyancy the fuel gases attain a velocity yb in the flame zone after leavingthe port exit, continuity requires that the effective radial diffusion distance be somevalue rb. Obviously, continuity requires that r be the same for both cases, so that:

r2j y ¼ r2byb

Thus, one observes that regardless of whether the fuel jet is momentum- orbuoyancy-controlled, the flame height yF is directly proportional to the volumetricflow leaving the port exit.

Given the condition that buoyancy can play a significant role, the fuel gases startwith an axial velocity and continue with a mean upward acceleration g due to buoy-ancy. The velocity of the fuel gases y is then given by:

y ¼ �y20 þ y2b�1=2

(6.25)

where y is the actual velocity, y0 is the momentum-driven velocity at the port, and ybis the velocity due to buoyancy. However, the buoyancy term can be closely approx-imated by:

y2b ¼ 2gyF (6.26)

where g is the acceleration due to buoyancy. If one substitutes Eqn (6.26) intoEqn (6.25) and expands the result in terms of a binomial expression, one obtains:

y ¼ y0

"1þ

gyF

y20

� 1

2

gyF

y20

2

þ/

#(6.27)

where the term in parentheses is the inverse of the modified Froude number:

Frh

y20gyF

(6.28)

316 CHAPTER 6 Diffusion flames

Page 17: Combustion || Diffusion flames

Thus, Eqn (6.27) can be written as:

y ¼ y0

�1þ

1

Fr

1

2Fr2

þ/

�(6.29)

For large Froude numbers, the diffusion flame height is momentum-controlledand y¼ y0. However, most laminar burning fuel jets will have very small Froudenumbers and y¼ yb; that is, most laminar fuel jets are buoyancy-controlled.

Although the flame height is proportional to the fuel volumetric flow ratewhether the flame is momentum- or buoyancy-controlled, the time to the flametip does depend on what the controlling force is. The characteristic time for diffusion(tD) must be equal to the time (ts) for a fluid element to travel from the port to theflame tip; that is, if the flame is momentum-controlled:

tDw

r2jD

!¼ yFy

�wts (6.30)

It follows from Eqn (6.30), of course, that:

yFwr2j y

Dw

pr2j y

pDw

Q

pD(6.31)

Equation (6.31) shows the same dependence on Q as that developed from theBurkeeSchumann approach (Eqns (6.21)e(6.23)). For a momentum-controlledfuel jet flame, the diffusion distance is rj, the jet port radius; and from Eqn (6.30)it is obvious that the time to the flame tip is independent of the fuel volumetricflow rate. For a buoyancy-controlled flame, ts remains proportional to (yF/y); how-ever, since y¼ (2gyF)

1/2:

twyFyw

yF�yF�1=2wy

1=2F wQ1=2 (6.32)

Thus, the stay time of a fuel element in a buoyancy-controlled laminar diffusionflame is proportional to the square root of the fuel volumetric flow rate. Thisconclusion is significant with respect to the soot smoke height tests to be discussedin Chapter 8.

For low Froude number, a fuel leaving a circular jet port immediately becomesaffected by a buoyancy condition, and the flame shape is determined by thiscondition and the consumptionof the fuel. In essence, followingEqn (6.26) the velocitywithin the exit fuel jet becomes proportional to its height above the port exit as theflow becomes buoyantly controlled. For example, the velocity along the axis wouldvary according to the proportionality y w y�1/2. Thus, the velocity increases as it ap-proaches the flame tip. Conservation of mass requires _m ¼ ryA, where _m is the massflow rate and r the density of the flowing gaseous fuel (or the fuel plus additives). Thusthe flame shape contracts not only due to the consumption of the fuel but also as thebuoyant velocity increases and a conical flame shape develops just above the port exit.

6.2 Gaseous Fuel Jets 317

Page 18: Combustion || Diffusion flames

For a constant mass flow buoyant system yw yF1/2 w QF

1/2. Thus yw P�1/2, and sincethe mass flow must be constant rather than affected by pressure change, _m ¼ ryA wPP�1/2P�1/2 ¼ constant. This proportionality states that, as the pressure of an exper-iment is raised, the planar cross-sectional area across any part of a diffusion flamemust contract and take a narrower conical shape. This pressure effect has been clearlyshown experimentally by Flowers and Bowman [12].

The preceding analyses hold only for circular fuel jets. Roper [10] has shown,and the experimental evidence verifies [11], that the flame height for a slot burneris not the same for momentum- and buoyancy-controlled jets. Consider a slot burnerof the WolfhardeParker type in which the slot width is x and the length is L. Asdiscussed earlier for a buoyancy-controlled situation, the diffusive distance wouldnot be x, but some smaller width, say xb. Following the terminology ofEqn (6.25), for a momentum-controlled slot burner:

twx2

Dw

yFy0

(6.33)

and:

yFwy0x

2=D

L=Lw

Q

D

xbL

�(6.34)

For a buoyancy-controlled slot burner:

twðxbÞ2D

wyFyb

(6.35)

yFwybðxbÞ2=D

L=Lw

Q

D

xbL

�(6.36)

Recalling that xb must be a function of the buoyancy, one has:

xbLybwQ

xbwQ

ybLw

Q�2gyF

�1=2L

Thus Eqn (6.36) becomes:

yFw

Q2

DL2ð2gÞ1=2

!2=3

w

Q4

D2L42g

1=3

(6.37)

Comparing Eqns (6.34) and (6.37), one notes that under momentum-controlledconditions for a given Q, the flame height is directly proportional to the slot width,while that under buoyancy-controlled conditions for a given Q, the flame height isindependent of the slot width. Roper et al. [11] have verified these conclusionsexperimentally.

318 CHAPTER 6 Diffusion flames

Page 19: Combustion || Diffusion flames

6.2.5 CONSERVED SCALARS AND MIXTURE FRACTIONThe simple coupling function U introduced in the last two sections is an exampleof a conserved scalar, which is a property that is conserved throughout the flowfield being neither created nor destroyed by chemical reaction. As shown in theprevious section, the classical approach to solving diffusion flame problems is todescribe the mixing by obtaining a solution for a conserved scalar. The assumptionof fast chemistry implies that the species concentrations and temperature arefunctions only of the conserved scalar. There are many scalar variables that areconserved. When the chemistry is simplified to a one-step reaction, simpleconserved scalars may be used, as was done in the previous sections. If theone-step reaction is written in terms of mass quantities instead of molar coeffi-cients as:

Fþ rO/ðr þ 1ÞPwhere MWFnF¼ 1 and MWOnO¼ r, and recognizing that:

_mF ¼ _mO

r¼ _mP

r þ 1

the following conserved scalars are attained among the fuel, oxidizer, and product,which are examples of coupling functions [13]:

UFO hmF � mO

r

UFP hmF þ mP

ðr þ 1Þ

UOP hmO þ rmP

ðr þ 1ÞAs shown previously, coupling functions can also be formed using the sensible

enthalpy. Simple conserved scalars are all linearly related so that the solution forone yields all of the others.

For a one-step reaction, the mixture fraction (Z), also a conserved scalar, can bedefined in terms of the fuel oxidizer coupling function having a value of 1 in the fuelstream and a value of 0 in the oxidizer stream:

Z ¼ U� UO;0

UF;0 � UO;0

where UF,0 and UO,0 are the values of U (¼ aF� aO) evaluated in the fuel andoxidizer streams (e.g., at the entrance boundary or infinity), respectively.

In systems that are more complex, one particular conserved scalar may befavored over another, for example, in systems with non-equal diffusivity, systemswith more than one feed for the fuel or oxidizer, or systems with complex chem-istry. With multistep reaction mechanisms, conserved scalars based on the

6.2 Gaseous Fuel Jets 319

Page 20: Combustion || Diffusion flames

species elements are generally used. The local mass fraction of element j may bedefined as:

Zj ¼XNi¼1

MWj

MWi

vðjÞi mi

whereMWj is the atomic weight of element j,MWi the molecular weight of species i,and v

ðjÞi the number of atoms of type j in species i. With the mixture fraction defined

as the mass of material having its origin in the fuel stream relative to the mass of themixture, the mixture fraction may be defined as:

Z ¼ Zj � Zj;O;0Zj;F;0 � Zj;O;0

where the element j does not have the same value of Zj in each stream. Again, Z isunity in the fuel stream and zero in the oxidizer stream. Using the mixture fractioncoordinate versus x or r, profiles of temperature and species concentration are muchless strongly coupled on the flow configuration [7,14]. The mixture fraction is animportant descriptor of diffusion flames and has been shown useful in describingthe structure of turbulent flames [7,14].

6.2.6 TURBULENT FUEL JETSSection 6.2.4 considered the burning of a laminar fuel jet, and the essential resultwith respect to flame height was that:

yF;Lw

r2j y

D

!w

Q

D

(6.38)

where yF,L specifies the flame height. When the fuel jet velocity is increased tothe extent that the flow becomes turbulent, the Froude number becomes largeand the system becomes momentum-controlleddmoreover, molecular diffusionconsiderations lose their validity under turbulent fuel jet conditions. One wouldintuitively expect a change in the character of the flame and its height. Turbulentflows are affected by the occurrence of flames, as discussed for premixedflame conditions, and they are affected by diffusion flames by many of the samemechanisms discussed for premixed flames. However, the diffusion flame in aturbulent mixing layer between the fuel and oxidizer streams steepens themaximum gradient of the mean velocity profile somewhat and generates vorticityof opposite signs on opposite sides of the high-temperature, low-density reactionregion.

The decreased overall density of the mixing layer with combustion increases thedimensions of the large vortices and reduces the rate of entrainment of fluids into themixing layer [14]. Thus it is appropriate to modify the simple phenomenologicalapproach that led to Eqn (6.31) to account for turbulent diffusion by replacing the

320 CHAPTER 6 Diffusion flames

Page 21: Combustion || Diffusion flames

molecular diffusivity with a turbulent eddy diffusivity. Consequently, the turbulentform of Eqn (6.38) becomes:

yF;Tw

r2j y

ε

!

where yF,T is the flame height of a turbulent fuel jet. But εw lU0, where l is the in-tegral scale of turbulence, which is proportional to the tube diameter (or radius rj),and U0 is the intensity of turbulence, which is proportional to the mean flow velocityy along the axis. Thus one may assume that:

εwrjy (6.39)

Combining Eqn (6.39) and the preceding equation, one obtains:

yF;Twr2j y

rjywrj (6.40)

This expression reveals that the height of a turbulent diffusion flame is propor-tional to the port radius (or diameter) above, irrespective of the volumetric fuelflow rate or fuel velocity issuing from the burner! This important practical conclu-sion has been verified by many investigators.

The earliest verification of yF,T w rj was by Hawthorne et al. [15], who reportedtheir results as yF,T as a function of jet exit velocity from a fixed tube exit radius.Thus varying the exit velocity is the same as varying the Reynolds number. Theresults in Ref. [15] were represented by Linan and Williams [14] in the diagramduplicated here as Figure 6.10. This figure clearly shows that as the velocity

4000

40

80

Liftoff

Blowoff

TurbulentLaminar

Hei

ght (

cm)

Exit velocity (m/s)

FIGURE 6.10

Variation of the character (height) of a gaseous diffusion flame as a function of fuel jet velocity

showing experimental flame liftoff.

After Linan and Williams [14] and Hawthorne et al. [15].

6.2 Gaseous Fuel Jets 321

Page 22: Combustion || Diffusion flames

increases in the laminar range, the height increases linearly, in accordance with Eqn(6.38). After transition to turbulence, the height becomes independent of the veloc-ity, in agreement with Eqn (6.40). Observing Figure 6.10, one notes that the transi-tion to turbulence begins near the top of the flame; then as the velocity increases, theturbulence rapidly recedes to the exit of the jet. At a high enough velocity, the flow inthe fuel tube becomes turbulent, and turbulence is observed everywhere in the flame.Depending on the fuel mixture, liftoff usually occurs after the flame becomes fullyturbulent. As the velocity increases further after liftoff, the liftoff height (the axialdistance between the fuel jet exit and the point where combustion begins) increasesapproximately linearly with the jet velocity [14]. After the liftoff height increases tosuch an extent that it reaches a value comparable to the flame diameter, a furtherincrease in the velocity causes blowoff [14].

6.3 BURNING OF CONDENSED PHASESWhen most liquids or solids are projected into an atmosphere so that a combustiblemixture is formed, an ignition of this mixture produces a flame that surrounds theliquid or solid phase. Except at the very lowest of pressures, around 10�6 Torr,this flame is a diffusion flame. If the condensed phase is a liquid fuel and the gaseousoxidizer is oxygen, the fuel evaporates from the liquid surface and diffuses to theflame front as the oxygen moves from the surroundings to the burning front. Thispicture of condensed-phase burning is most readily applied to droplet burning, butcan also be applied to any liquid surface.

The rate at which the droplet evaporates and burns is generally considered to bedetermined by the rate of heat transfer from the flame front to the fuel surface. Here,as in the case of gaseous diffusion flames, chemical processes are assumed to occurso rapidly that the burning rates are determined solely by mass and heat transferrates.

Many of the early analytical models of this burning process considered a double-film model for the combustion of the liquid fuel. One film separated the dropletsurface from the flame front and the other separated the flame front from thesurrounding oxidizer atmosphere, as depicted in Figure 6.11.

In some analytical developments the liquid surface was assumed to be at thenormal boiling point of the fuel. Surveys of the temperature field in burning liquidsby Khudyakov [16] indicated that the temperature is just a few degrees below theboiling point. In the approach to be employed here, the only requirement is thatthe droplet be at a uniform temperature at or below the normal boiling point.In the sf region of Figure 6.11, fuel evaporates at the drop surface and diffuses to-ward the flame front where it is consumed. Heat is conducted from the flame frontto the liquid and vaporizes the fuel. Many analyses assume that the fuel is heated tothe flame temperature before it chemically reacts and that the fuel does not reactuntil it reaches the flame front. This latter assumption implies that the flame frontis a mathematically thin surface where the fuel and oxidizer meet in stoichiometric

322 CHAPTER 6 Diffusion flames

Page 23: Combustion || Diffusion flames

proportions. Some early investigators first determined Tf in order to calculate thefuel burning rate. However, in order to determine a Tf, the infinitely thin reactionzone at the stoichiometric position must be assumed.

In the film fN, oxygen diffuses to the flame front, and combustion products andheat are transported to the surrounding atmosphere. The position of the boundarydesignated by N is determined by convection. A stagnant atmosphere places theboundary at an infinite distance from the fuel surface.

Although most analyses assume no radiant energy transfer, as will be shown sub-sequently, the addition of radiation poses no mathematical difficulty in the solutionto the mass burning rate problem.

6.3.1 GENERAL MASS BURNING CONSIDERATIONS AND THEEVAPORATION COEFFICIENT

Three parameters are generally evaluated: the mass burning rate (evaporation), theflame position above the fuel surface, and the flame temperature. The most impor-tant parameter is the mass burning rate, for it permits the evaluation of the so-calledevaporation coefficient, which is most readily measured experimentally.

The use of the term evaporation coefficient comes about from mass and heattransfer experiments without combustiond that is, evaporation, as generally usedin spray-drying and humidification. Basically, the evaporation coefficient b isdefined by the following expression, which has been verified experimentally:

d2 h d20 � bt (6.41)

fs0

0.5

1.0

Droplet flame radius

Fuel

Products

T�

T��TTf

Nitrogen

Oxidizer

mA � ρρA

FIGURE 6.11

Characteristic parametric variations of dimensionless temperature, T 0, and mass fraction

m of fuel, oxygen, and products along a radius of a droplet diffusion flame in a quiescent

atmosphere. Tf is the adiabatic, stoichiometric flame temperature, rA is the partial density of

species A, and r is the total mass density. The estimated values derived for benzene are given

in Section 6.3.2.2.

6.3 Burning of Condensed Phases 323

Page 24: Combustion || Diffusion flames

where d0 is the original drop diameter and d is the drop diameter after time t. It willbe shown later that the same expression must hold for mass and heat transfer withchemical reaction (combustion).

The combustion of droplets is one aspect of a much broader problem, which in-volves the gasification of a condensed phase, that is, a liquid or a solid. In this sense,the field of diffusion flames is rightly broken down into gases and condensed phases.Here the concern is with the burning of droplets, but the concepts to be used are justas applicable to other practical experimental properties such as the evaporation of liq-uids, sublimation of solids, hybrid burning rates, ablation heat transfer, solid propel-lant burning, transpiration cooling, and the like. In all categories, the interest is in themass consumption rate, or the rate of regression of a condensed-phase material. Ingaseous diffusion flames, there was no specific property to measure and the flameheight was evaluated; but in condensed-phase diffusion flames, a specific quantityis measurable. This quantity is some representation of the mass consumption rateof the condensed phase. The similarity to the case just mentioned arises from thefact that the condensed phase must be gasified; consequently, there must be an energyinput into the condensed material. What determines the rate of regression or evolu-tion of material is the heat flux at the surface. Thus, in all the processes mentioned:

q ¼ _rrfL0v (6.42)

where q is the heat flux to the surface in calories per square centimeter per second; _ris the regression rate in centimeters per second; rf is the density of the condensedphase; and L0v is the overall energy per unit mass required to gasify the material.Usually, L0v is the sum of two termsdthe heat of vaporization, sublimation, or gasi-fication plus the enthalpy required to bring the surface to the temperature of vapor-ization, sublimation, or gasification.

From the foregoing discussion, it is seen that the heat flux q, L0v, and the densitydetermine the regression rate; but this statement does not mean that the heat flux isthe controlling or rate-determining step in each case. In fact, it is generally not thecontrolling step. The controlling step and the heat flux are always interrelated, how-ever. Regardless of the process of concern (assuming no radiation):

q ¼ �l

vT

vy

s

(6.43)

where l is the thermal conductivity and the subscript “s” designates the fuel surface.This simple statement of the Fourier heat conduction law is of such great signifi-cance that its importance cannot be overstated.

This same equation holds whether or not there is mass evolution from the surfaceand whether or not convective effects prevail in the gaseous stream. Even for convec-tive atmospheres in which one is interested in the heat transfer to a surface (withoutmass addition of any kind, that is, the heat transfer situation generally encountered),one writes the heat transfer equation as:

q ¼ hðTN � TsÞ (6.44)

324 CHAPTER 6 Diffusion flames

Page 25: Combustion || Diffusion flames

Obviously, this statement is shorthand for:

q ¼ �l

vT

vy

s

¼ hðTN � TsÞ (6.45)

where TN and Ts are the free-stream and surface temperatures, respectively; the heattransfer coefficient h is by definition:

hhl

d(6.46)

where d is the boundary layer thickness. Again by definition, the boundary layer isthe distance between the surface and free-stream condition; thus, as anapproximation:

q ¼ lðTN � TsÞd

(6.47)

The (TN� Ts)/d term is the temperature gradient, which correlates (vT/vy)sthrough the boundary layer thickness. The fact that d can be correlated with the Rey-nolds number and that the Colburn analogy can be applied leads to the correlation ofthe form:

Nu ¼ f ðRe;PrÞ (6.48)

where Nu is the Nusselt number (hx/l), Pr the Prandtl number (cpm/l), and Re theReynolds number (ryx/m); here x is the critical dimensiondthe distance from theleading edge of a flat plate or the diameter of a tube.

Although the correlations given by Eqn (6.48) are useful for practical evalua-tion of heat transfer to a wall, one must not lose sight of the fact that the temper-ature gradient at the wall actually determines the heat flux there. In transpirationcooling problems, it is not so much that the injection of the transpiring fluid in-creases the boundary layer thickness, thereby decreasing the heat flux, but ratherthat the temperature gradient at the surface is decreased by the heat-absorbingcapability of the injected fluid. What Eqn (6.43) specifies is that regardless ofthe processes taking place, the temperature profile at the surface determines theregression ratedwhether it be evaporation, solid propellant burning, etc. Thus,all the mathematical approaches used in this type of problem simply seek to eval-uate the temperature gradient at the surface. The temperature gradient at thesurface is different for the various processes discussed. Thus, the temperature pro-file from the surface to the source of energy for evaporation will differ from that forthe burning of a liquid fuel, which releases energy when oxidized in a flamestructure.

Nevertheless, a diffusion mechanism generally prevails; and because it is theslowest step, it determines the regression rate. In evaporation, the mechanism isthe conduction of heat from the surrounding atmosphere to the surface; in ablation,it is the conduction of heat through the boundary layer; in droplet burning, it is therates at which the fuel diffuses to approach the oxidizer, etc.

6.3 Burning of Condensed Phases 325

Page 26: Combustion || Diffusion flames

It is mathematically interesting that the gradient at the surface will always be aboundary condition to the mathematical statement of the problem. Thus, the math-ematical solution is necessary simply to evaluate the boundary condition.

Furthermore, it should be emphasized that the absence of radiation has beenassumed. Incorporating radiation transfer is not difficult if one assumes that theradiant intensity of the emitters is known and that no absorption occurs betweenthe emitters and the vaporizing surfaces, that is, it can be assumed that qr, the radiantheat flux to the surface, is known. Then Eqn (6.43) becomes:

qþ qr ¼ l

vT

vy

s

þ qr ¼ rrfL0v (6.49)

Note that using these assumptions does not make the mathematical solutionof the problem significantly more difficult, for again, qrdand hence radiationtransferdis a known constant and enters only in the boundary condition. The differ-ential equations describing the processes are not altered.

First, the evaporation rate of a single fuel droplet is calculated before consideringthe combustion of this fuel droplet; or, to say it more exactly, one calculates theevaporation of a fuel droplet in the absence of combustion. Since the concern iswith diffusional processes, it is best to start by reconsidering the laws that holdfor diffusional processes.

Fick’s law states that if a gradient in concentration of speciesA exists, say (dnA/dy),a flow or flux of A, say jA, across a unit area in the y direction will be proportional tothe gradient so that:

jA ¼ �DdnAdy

(6.50)

where D is the proportionality constant called the molecular diffusion coefficient or,more simply, the diffusion coefficient; nA is the number concentration of moleculesper cubic centimeter; and j is the flux of molecules, in number of molecules persquare centimeter per second. Thus, the unit of D is square centimeters per second.

The Fourier law of heat conduction relates the flux of heat q per unit area, as aresult of a temperature gradient, such that:

q ¼ �ldT

dy

The unit of q is calories per square centimeter per second, and the unit of the ther-mal conductivity l is calories per centimeter per second per degree Kelvin. It is notthe temperature, an intensive thermodynamic property, that is exchanged, but energycontent, an extensive property. In this case, the energy density and the exchange re-action, which show similarity, are written as:

q ¼ � l

rcprcp

dT

dy¼ � l

rcp

d�rcpT

�dy

¼ �adH

dy(6.51)

326 CHAPTER 6 Diffusion flames

Page 27: Combustion || Diffusion flames

where a is the thermal diffusivity, the unit of which is square centimeters per secondsince l¼ cal/cm-s-K, cp¼ cal/g-K, and r¼ g/cm3; and H is the energy concentra-tion in calories per cubic centimeter. Thus, the similarity of Fick’s and Fourier’slaws is apparent. The former is due to a number concentration gradient and the latterto an energy concentration gradient.

A law similar to these two diffusional processes is Newton’s law of viscosity,which relates the flux (or shear stress) syx of the x component of momentum dueto a gradient in ux; this law is written as:

syx ¼ �mduxdy

(6.52)

where the units of the stress s are dynes per square centimeter and those of the vis-cosity are grams per centimeter per second. Again, it is not velocity that isexchanged, but momentum; thus when the exchange of momentum density is writ-ten, similarity is again noted:

syx ¼ �m

r

�dðruxÞdy

�¼ �n

�dðruxÞdy

�(6.53)

where n is the momentum diffusion coefficient or, more acceptably, the kinematicviscosity; n is a diffusivity and its unit is also square centimeters per second. SinceEqn (6.53) relates the momentum gradient to a flux, its similarity to Eqns (6.50) and(6.51) is obvious. Recall, as stated in Chapter 4, that the simple kinetic theory forgases predicts a¼D¼ n. The ratios of these three diffusivities give some familiardimensionless similarity parameters:

Pr ¼ n

a; Sc ¼ n

D; Le ¼ a

D

where Pr, Sc, and Le are Prandtl, Schmidt, and Lewis numbers, respectively. Thus,for gases simple kinetic theory gives as a first approximation:

Pr ¼ Sc ¼ Le ¼ 1

6.3.2 SINGLE FUEL DROPLETS IN QUIESCENT ATMOSPHERESSince Fick’s law will be used in many different forms in the ensuing development, itis best to develop those forms so that the later developments need not be interrupted.

Consider the diffusion of molecules A into an atmosphere of B molecules, that is,a binary system. For a concentration gradient in A molecules alone, future develop-ments can be simplified readily if Fick’s law is now written:

jA ¼ �DABdnAdy

(6.54)

where DAB is the binary diffusion coefficient. Here, nA is considered in number ofmoles per unit volume, since one could always multiply Eqn (6.50) through byAvogadro’s number, and jA is expressed in number of moles as well.

6.3 Burning of Condensed Phases 327

Page 28: Combustion || Diffusion flames

Multiplying through Eqn (6.54) byMWA, the molecular weight of A, one obtains:�jAMWA

� ¼ �DABdðnAMWAÞ

dy¼ �DAB

drAdy

(6.55)

where rA is the mass density of A, rB is the mass density of B, and n is the total num-ber of moles:

n ¼ nA þ nB (6.56)

which is constant, and:

dn

dy¼ 0;

dnAdy

¼ �dnBdy

; jA ¼ �jB (6.57)

The result is a net flux of mass by diffusion equal to:

ry ¼ jAMWA þ jBMWB (6.58)

¼ jAðMWA �MWBÞ (6.59)

where y is the bulk direction velocity established by diffusion.In problems dealing with the combustion of condensed matter, and hence

regressing surfaces, there is always a bulk velocity movement in the gases. Thus,species are diffusing while the bulk gases are moving at a finite velocity. The diffu-sion of the species can be against or with the bulk flow (velocity). For mathematicalconvenience, it is best to decompose such flows into a flow with an average massvelocity y and a diffusive velocity relative to y.

When one gas diffuses into another, as A into B, even without the quasi-steady-flow component imposed by the burning, the mass transport of a species, say A, ismade up of two componentsdthe normal diffusion component and the componentrelated to the bulk movement established by the diffusion process. This mass trans-port flow has a velocity DA and the mass of A transported per unit area is rADA. Thebulk velocity established by the diffusive flow is given by Eqn (6.58). The fraction ofthat flow is Eqn (6.58) multiplied by the mass fraction of A, rA/r. Thus:

rADA ¼ �DAB

drAdy

�1��1�MWB

MWA

�rA

r

�(6.60)

Since jA¼�jB:

MWAjA ¼ �MWAjB ¼ ��MWBjB�MWA

MWB

(6.61)

and:

�drAdy

¼MWA

MWB

drBdy

(6.62)

However:

drAdy

þ drBdy

¼ dr

dy(6.63)

328 CHAPTER 6 Diffusion flames

Page 29: Combustion || Diffusion flames

which gives with Eqn (6.62):drAdy

�MWB

MWA

drAdy

¼ dr

dy(6.64)

Multiplying through by rA/r, one obtains:rA

r

�1�MWB

MWA

�drAdy

¼rA

r

dr

dy

(6.65)

Substituting Eqn (6.65) into Eqn (6.60), one finds:

rADA ¼ �DAB

"drAdy

�rA

r

dr

dy

#(6.66)

or:

rADA ¼ �rDABdðrA=rÞ

dy(6.67)

Defining mA as the mass fraction of A, one obtains the following proper form forthe diffusion of species A in terms of mass fraction:

rADA ¼ �rDABdmA

dy(6.68)

This form is that most commonly used in the conservation equation.The total mass flux of A under the condition of the burning of a condensed phase,

which imposes a bulk velocity developed from the mass burned, is then:

rAyA ¼ rAyþ rADA ¼ rAy� rDABdmA

dy(6.69)

where rAy is the bulk transport part and rADA is the diffusive transport part. Indeed,in the developments of Chapter 4, the correct diffusion term was used without theproof just completed.

6.3.2.1 Heat and mass transfer without chemical reaction (evaporation):the transfer number B

Following Blackshear’s [17] adaptation of Spalding’s approach [18,19], consider-ation will now be given to the calculation of the evaporation of a single fuel dropletin a nonconvective atmosphere at a given temperature and pressure. A majorassumption is now made in that the problem is considered as a quasi-steady one.Thus, the droplet is of fixed size and retains that size by a steady flux of fuel. Onecan consider the regression as being constant; or even better, one can think ofthe droplet as a porous sphere being fed from a very thin tube at a rate equal tothe mass evaporation rate so that the surface of the sphere is always wet and anyliquid evaporated is immediately replaced. The porous sphere approach shows

6.3 Burning of Condensed Phases 329

Page 30: Combustion || Diffusion flames

that for the diffusion flame, a bulk gaseous velocity outward must exist; andalthough this velocity in the spherical geometry will vary radially, it must alwaysbe the value given by _m ¼ 4pr2ry. This velocity is the one referred to in the last sec-tion. With this physical picture one may take the temperature throughout the dropletas constant and equal to the surface temperature as a consequence of the quasi-steady assumption.

In developing the problem, a differential volume in the vapor above the liquiddroplet is chosen, as shown in Figure 6.12. The surface area of a sphere is 4pr2.Since mass cannot accumulate in the element:

dðrAyÞ ¼ 0; ðd=drÞ�4pr2ry� ¼ 0 (6.70)

which is essentially the continuity equation. Consider now the energy equation ofthe evaporating droplet in spherical-symmetric coordinates in which cp and l aretaken independent of temperature. The heat entering at the surface (i.e., the amountof heat convected in) is _mcpT or (4pr2ry)cpT (see Figure 6.12). The heat leavingafter rþDr is _mcp½T þ ðdT=drÞDr� or ð4pr2ryÞcp½T þ ðdT=drÞDr�. The difference,then, is:

�4pr2rycpðdT=drÞDrThe heat diffusing from r toward the drop (out of the element) is:

�l4pr2ðdT=drÞThe heat diffusing into the element is:

�l4pðr þ DrÞ2ðd=drÞ½T þ ðdT=drÞDr�or:

�"l4pr2

dT

dr

þ l4pr2

d2T

dr2

Dr þ l8prDr

dT

dr

#

T

r

Δr

r � Δr

T �dTdr

�r

FIGURE 6.12

Temperature balance across a differential element of a diffusion flame in spherical symmetry.

330 CHAPTER 6 Diffusion flames

Page 31: Combustion || Diffusion flames

plus two terms in Dr2 and one in Dr3 which are negligible. The difference in thetwo terms is:

þ�l4pr2�d2T�dr2�Dr þ 8lprðdT=drÞDr�Heat could be generated in the volume element defined by Dr, so one has:

þ�4pr2Dr� _Hwhere _H is the rate of enthalpy change per unit volume. Thus, for the energybalance:

4pr2rycpðdT=drÞDr � l4pr2�d2T

�dr2�Dr � l8prðdT=drÞDr ¼

þ4pr2Dr _H(6.71)

4pr2rycp

dT

dr

¼ l4pr2

d2T

dr2

þ 8lpr

dT

dr

þ 4pr2 _H (6.72)

or:

4pr2�ry�dcpT

dr

¼ d

dr

"l4pr2

cp

dcpT

dr

þ 4pr2 _H

#(6.73)

Similarly, the conservation of the Ath species can be written as:

4pr2ry

dmA

dr

¼d

dr

�4pr2rD

dmA

dr

�þ 4pr2 _mA (6.74)

where _mA is the generation or disappearance rate of A due to reaction in the unitvolume. According to the kinetic theory of gases, to a first approximation the prod-uct rD (and hence l/cp) is independent of temperature and pressure; consequently,rsDs¼ rD, where the subscript “s” designates the condition at the droplet surface.

Consider a droplet of radius r. If the droplet is vaporizing, the fluid will leave thesurface by convection and diffusion. Since at the liquid droplet surface only A exists,the boundary condition at the surface is:

rl _r ¼ rsys|fflfflfflfflfflffl{zfflfflfflfflfflffl}amount of materialleaving the surface

¼ rmAsys � rD

dmA

dr

s

(6.75)

where rl is the liquid droplet density and _r is the rate of change of the liquid dropletradius. Equation (6.75) is, of course, explicitly Eqn (6.69) when rsys is the bulk massmovement, which at the surface is exactly the amount of A that is being convected(evaporated) written in terms of a gaseous density and velocity plus the amount ofgaseous A that diffuses to or from the surface. Since products and inerts diffuse tothe surface, mAs has a value less than 1. Equation (6.75), then, is written as:

ys ¼ DðdmA=drÞsðmAs � 1Þ (6.76)

6.3 Burning of Condensed Phases 331

Page 32: Combustion || Diffusion flames

In the sense of Spalding, a new parameter b is defined:

bhmA

mAs � 1

Equation (6.76) thus becomes:

ys ¼ D

db

dr

s

(6.77)

and Eqn (6.74) written in terms of the new variable b and for the evaporationcondition ði:e:; _mA ¼ 0Þ is:

r2ry

db

dr

¼d

dr

�r2rD

db

dr

�(6.78)

The boundary condition at r¼N is mA¼mAN or:

b ¼ bN at r/N (6.79)

From continuity:

r2ry ¼ r2s rsys (6.80)

Since r2ry¼ constant on the left-hand side of the equation, integration of Eqn(6.78) proceeds simply and yields:

r2ryb ¼ r2rD

db

dr

þ const (6.81)

Evaluating the constant at r¼ rs, one obtains:

r2s rsysbs ¼ r2s rsys þ const

since from Eqn (6.77) ys¼D(db/dr)s. Or, one has from Eqn (6.81):

r2s rsysðb� bs þ 1Þ ¼ r2rD

db

dr

(6.82)

By separating variables: r2s rsysr2rD

dr ¼ db

b� bs þ 1(6.83)

Assuming rD is constant, and integrating (recall that rD¼ rsDs), one has:

�r2s ysrDs

¼ lnðb� bs þ 1Þ þ const (6.84)

Evaluating the constant at r/N, one obtains:

const ¼ �lnðbN � bs þ 1Þ

332 CHAPTER 6 Diffusion flames

Page 33: Combustion || Diffusion flames

or Eqn (6.84) becomes:

r2srDs

¼ ln

�bN � bs þ 1

b� bs þ 1

�(6.85)

The left-hand term of Eqn (6.84) goes to zero as r/N. This point is significantbecause it shows that the quiescent spherical-symmetric case is the only mathemat-ical case that does not blow up. No other quiescent case, such as that for cylindricalsymmetry or any other symmetry, is tractable. Evaluating Eqn (6.85) at r¼ rs resultsin:

rsysDs

¼ lnðbN � bs þ 1Þ ¼ lnð1þ BÞ

rsys ¼ Ds lnðbN � bs þ 1Þ ¼ Ds lnð1þ BÞ

rsys ¼ Ds ln

�1þ

�mAN � mAs

mAs � 1

�� (6.86)

Here Bh bN� bs and is generally referred to as the Spalding transfer number.The mass consumption rate per unit area GA ¼ _mA=4pr

2s , where _mA ¼ rsys4pr

2s , is

then found by multiplying Eqn (6.86) by 4pr2s rs=4pr2s and cross-multiplying rs, to

give:

4pr2s rsys

4pr2s¼ rsDs

rslnð1þ BÞ

GA ¼ _mA

4pr2s¼rsDs

rs

lnð1þ BÞ ¼

rD

rs

lnð1þ BÞ

(6.87)

Since the product rD is independent of pressure, the evaporation rate is essen-tially independent of pressure. There is a mild effect of pressure on the transfer num-ber, as will be discussed in more detail when the droplet burning case is considered.In order to find a solution for Eqn (6.87) or, more rightly, to evaluate the transfernumber B, mAs must be determined. A reasonable assumption would be that thegas surrounding the droplet surface is saturated at the surface temperature Ts. Sincevapor pressure data are available, the problem then is to determine Ts.

For the case of evaporation, Eqn (6.73) becomes:

r2s rsyscpdT

dr¼ d

dr

�r2l

dT

dr

�(6.88)

which, upon integration, becomes:

r2s rsyscpT ¼ r2l

dT

dr

þ const (6.89)

6.3 Burning of Condensed Phases 333

Page 34: Combustion || Diffusion flames

The boundary condition at the surface is:�l

dT

dr

�s

¼ rsysLv (6.90)

where Lv is the latent heat of vaporization at the temperature Ts. Recall that thedroplet is considered uniform throughout at the temperature Ts. Substituting Eqn(6.90) into Eqn (6.89), one obtains:

const ¼ r2s rsys�cpTs � Lv

�Thus, Eqn (6.89) becomes:

r2s rsyscp

�T � Ts þ

Lvcp

�¼ r2l

dT

dr

(6.91)

Integrating Eqn (6.91), one has:

�r2s rsyscprl

¼ ln

�T � Ts þ

Lvcp

�þ const (6.92)

After evaluating the constant at r/N, one obtains:

r2s rsyscprl

¼ ln

"TN � Ts þ

�Lv�cp�

T � Ts þ�Lv�cp� # (6.93)

Evaluating Eqn (6.93) at the surface (r¼ rs; T¼ Ts) gives:

rsrsyscpl

¼ ln

cpðTN � TsÞ

Lvþ 1

(6.94)

And since a¼ l/rcp:

rsys ¼ as ln

1þ cpðTN � TsÞ

Lv

(6.95)

Comparing Eqns (6.86) and (6.95), one can write:

rsys ¼ as ln

cpðTN � TsÞ

Lvþ 1

¼ Ds ln

�1þ

mAN � mAs

mAs � 1

�or:

rsys ¼ as lnð1þ BTÞ ¼ Ds lnð1þ BMÞ (6.96)

where:

BThcp�TN � Ts

�Lv

; BMhmAN � mAs

mAs � 1

Again, since a¼D:

BT ¼ BM

334 CHAPTER 6 Diffusion flames

Page 35: Combustion || Diffusion flames

and:

cp�TN � Ts

�Lv

¼ mAN � mAs

mAs � 1(6.97)

Although mAs is determined from the vapor pressure of A or the fuel, one mustrealize that mAs is a function of the total pressure, since:

mAs hrAs

r¼ nAMWA

nMW¼PA

P

MWA

MW

(6.98)

where nA and n are the number of moles of A and the total number of moles, respec-tively; MWA and MW are the molecular weight of A and the average molecularweight of the mixture, respectively; and PA and P are the vapor pressure of A andthe total pressure, respectively.

In order to obtain the solution desired, a value of Ts is assumed, the vapor pres-sure of A is determined from tables, and mAs is calculated from Eqn (6.98). Thisvalue of mAs and the assumed value of Ts are inserted in Eqn (6.97). If this equationis satisfied, the correct Ts is chosen. If not, one must reiterate. When the correct valueof Ts and mAs are found, BT or BM are determined for the given initial conditions TNor mAN. For fuel combustion problems, mAN is usually zero; however, for evapora-tion, say of water, there is humidity in the atmosphere, and this humidity must berepresented as mAN. Once BT and BM are determined, the mass evaporation rateis determined from Eqn (6.87) for a fixed droplet size. It is, or course, much prefer-able to know the evaporation coefficient b from which the total evaporation time canbe determined. Once B is known, the evaporation coefficient can be determinedreadily, as will be shown later.

6.3.2.2 Heat and mass transfer with chemical reaction (droplet burningrates)

The previous developments also can be used to determine the burning rate, or evap-oration coefficient, of a single droplet of fuel burning in a quiescent atmosphere. Inthis case, the mass fraction of the fuel, which is always considered to be thecondensed phase, will be designated mf, and the mass fraction of the oxidizermo. mo is the oxidant mass fraction exclusive of inerts, and i is used as the mass stoi-chiometric fueleoxidant ratio, also exclusive of inerts. The same assumptions thathold for evaporation from a porous sphere hold here. Recall that the temperaturethroughout the droplet is considered to be uniform and equal to the surface temper-ature Ts. This assumption is sometimes referred to as one of infinite condensed phasethermal conductivity. For liquid fuels, this temperature is generally near, but slightlyless than, the saturation temperature for the prevailing ambient pressure.

As in the case of burning gaseous fuel jets, it is assumed that the fuel and oxidantapproach each other in stoichiometric proportions. The stoichiometric relations arewritten as:

_mf ¼ _moi; _mfH ¼ _moHi ¼ � _H (6.99)

6.3 Burning of Condensed Phases 335

Page 36: Combustion || Diffusion flames

where _mf and _mo are the mass consumption rates per unit volume, H the heat of re-action (combustion) of the fuel per unit mass, and _H the heat release rate per unitvolume. The mass consumption rates _mf and _mo are decreasing.

There are now three fundamental diffusion equations: one for the fuel, onefor the oxidizer, and one for the heat. Equation (6.74) is then written as two equa-tions: one in terms of mf and the other in terms of mo. Equation (6.73) remainsthe same for evaporation and combustion. As seen in the case of the evaporation,the solution to the equations becomes quite simple when written in terms of theb variable, which led to the Spalding transfer number B. As noted in this case,the b variable was obtained from the boundary condition. Indeed, another b vari-able could have been obtained for the energy equation (Eqn (6.73))da variablehaving the form:

b ¼cpT

Lv

As in the case of burning gaseous fuel jets, the diffusion equations are combined

readily by assuming rD¼ (l/cp), that is, Le¼ 1. The same procedure can be fol-lowed in combining the boundary conditions for the three droplet burning equationsto determine the appropriate b variables to simplify the solution for the mass con-sumption rate.

The surface boundary condition for the diffusion of fuel is the same as that forpure evaporation (Eqn (6.75)) and takes the form:

rsys�mfs � 1

� ¼ rD

dmf

dr

s

(6.100)

Since there is no oxidizer leaving the surface, the surface boundary condition fordiffusion of oxidizer is:

0 ¼ rsysmos � rD

dmo

dr

s

or rsysmos ¼ rD

dmo

dr

s

(6.101)

The boundary condition for the energy equation is also the same as that wouldhave been for the droplet evaporation case (Eqn (6.75)), and is written as:

rsysLv ¼l

cp

�d�cpT�

dr

�s

(6.102)

Multiplying Eqn (6.100) by H and Eqn (6.101) by iH gives the new forms:

Hhrsys

�mfs � 1

�i ¼ �rDdmf

dr

s

�H (6.103)

and:

Hhi�rsysmos

�i ¼ ��rDdmo

dr

s

�i

�H (6.104)

336 CHAPTER 6 Diffusion flames

Page 37: Combustion || Diffusion flames

By adding Eqns (6.102) and (6.103) and recalling that rD¼ (l/cp), one obtains:

rsys

hH�mfs � 1

�þ Lv

i¼ rD

�d�mfH þ cpT

�dr

�s

and after transposing:

rsys ¼ rD

"d

dr

mfH þ cpT

Hðmfs � 1Þ þ Lv

#s

(6.105)

Similarly, by adding Eqns (6.102) and (6.104), one obtains:

rsys�Lv þ Himos

� ¼ rD

�d

dr

�cpT þ moiH

��s

or:

rsys ¼ rD

"d

dr

moiH þ cpT

Himos þ Lv

#s

(6.106)

And, finally, by subtracting Eqn (6.104) from Eqn (6.103), one obtains:

rsys

h�mfs � 1

�� imos

i¼ rD

�dðmf � imoÞ

dr

�s

or:

rsys ¼ rD

�d

dr

mf � imo

ðmfs � 1Þ � imos

�s

(6.107)

Thus, the new b variables are defined as:

bfqhmfH þ cpT

Hðmfs � 1Þ þ Lv; boqh

moiH þ cpT

Himos þ Lv;

bfohmf � imo

ðmfs � 1Þ � imos

(6.108)

The denominator of each of these three b variables is a constant. The three diffu-sion equations are transformed readily in terms of these variables by multiplying thefuel diffusion equation by H and the oxygen diffusion equation by iH. By using thestoichiometric relations (Eqn (6.99)) and combining the equations in the samemanner as the boundary conditions, one can eliminate the nonhomogeneous terms_mf ; _mo, and _H. Again, it is assumed that rD¼ (l/cp). The combined equations arethen divided by the appropriate denominators from the b variables so that all equa-tions become similar in form. Explicitly, then, one has the following developments:

r2ryd�cpT�

dr¼ d

dr

"r2

l

cp

d�cpT�

dr

#� r2 _H (6.109)

6.3 Burning of Condensed Phases 337

Page 38: Combustion || Diffusion flames

H

(r2ry

dmf

dr¼ d

dr

�r2rD

dmf

dr

�þ r2 _mf

)(6.110)

Hi

(r2ry

dmo

dr¼ d

dr

�r2rD

dmo

dr

�þ r2 _mo

)(6.111)

_mfH ¼ _moHi ¼ � _H (6.99)

Adding Eqns (6.109) and (6.110), dividing the resultant equation through by[H(mfs� 1)þ Lv], and recalling that _mfH ¼ � _H, one obtains:

r2ryd

dr

mfH þ cpT

Hðmfs � 1Þ þ Lv

¼ d

dr

"r2rD

d

dr

mfH þ cpT

Hðmfs � 1Þ þ Lv

#(6.112)

which is then written as:

r2ry

dbfqdr

¼ d

dr

�r2rD

dbfqdr

�(6.113)

Similarly, by adding Eqns (6.109) and (6.111) and dividing the resultant equationthrough by [Himosþ Lv], one obtains:

r2ryd

dr

moiH þ cpT

Himos þ Lv

¼ d

dr

"r2rD

d

dr

moiH þ cpT

Himos þ Lv

#(6.114)

or:

r2ryd

dr

�boq� ¼ d

dr

�r2rD

d

dr

�boq��

(6.115)

Following the same procedures by subtracting Eqn (6.111) from Eqn (6.110), oneobtains:

r2ryd

dr

�bfo� ¼ d

dr

�r2rD

d

drðbfoÞ

�(6.116)

Obviously, all the combined equations have the same form and boundary condi-tions, that is:

r2ryd

drðbÞfo;fq;oq ¼

d

dr

�r2rD

d

drðbÞfo;fq;oq

rsys ¼ rD

�d

drðbÞfo;fq;oq

�s

at r ¼ rs

b ¼ bN at r/N

(6.117)

338 CHAPTER 6 Diffusion flames

Page 39: Combustion || Diffusion flames

The equation and boundary conditions are the same as those obtained for the pureevaporation problem; consequently, the solution is the same. Thus, one writes:

Gf ¼ _mf

4pr2s¼rsDs

rs

ln�1þ B

�where B ¼ bN � bs (6.118)

It should be recognized that since Le¼ Sc¼ Pr, Eqn (6.118) can also be writtenas:

Gf ¼

l

cprs

lnð1þ BÞ ¼

m

rs

lnð1þ BÞ (6.119)

In Eqn (6.118), the transfer number B can take any of the following forms:

Without combustion assumption With combustion assumption

Bfo ¼ ðmfN � mfsÞ þ ðmos � moNÞiðmfs � 1Þ � imos

¼ ðimoN þ mfsÞ1� mfs

Bfq ¼H�mfN � mfs

�þ cp�TN � Ts

�Lv þ Hðmfs � 1Þ ¼ cp

�TN � Ts

�� mfsH

Lv þ Hðmfs � 1Þ

Boq ¼Hi�moN � mos

�þ cp�TN � Ts

�Lv þ imosH

¼ cp�TN � Ts

�þ imoNH

Lv

(6.120)

The combustion assumption in Eqn (6.120) is that mos¼mfN¼ 0, since it isassumed that neither fuel nor oxidizer can penetrate the flame zone. This require-ment is not that the flame zone be infinitely thin, but that all the oxidizer must beconsumed before it reaches the fuel surface and that the quiescent atmospherecontain no fuel.

As in the evaporation case, in order to solve Eqn (6.118), it is necessary to pro-ceed by first equating Bfo¼ Boq. This expression:

imoN þ mfs

1� mfs¼ cp

�TN � Ts

�þ imoNH

Lv(6.121)

is solved with the use of the vapor pressure data for the fuel. The iteration processdescribed in the solution of BM¼ BT in the evaporation problem is used. The solu-tion of Eqn (6.121) gives Ts and mfs and, thus, individually Bfo and Boq. With Bknown, the burning rate is obtained from Eqn (6.118).

For the combustion systems, Boq is the most convenient form of B: cp(TN� Ts),is usually much less than imoNH, and to a close approximation Boqy (imoNH/Lv).Thus the burning rate (and, as will be shown later, the evaporation coefficient b) isreadily determined. It is not necessary to solve for mfs and Ts. Furthermore, detailedcalculations reveal that Ts is very close to the saturation temperature (boiling point)of most liquids at the given pressure. Thus, although rD in the burning rate expres-sion is independent of pressure, the transfer number increases as the pressure israised because pressure rises concomitantly with the droplet saturation temperature.

6.3 Burning of Condensed Phases 339

Page 40: Combustion || Diffusion flames

As the saturation temperature increases, the latent heat of evaporation Lv decreases.Of course, Lv approaches zero at the critical point. This decrease in Lv is greater thanthe decrease of the cpg(TN� Ts) and cpl(Ts� Tl) terms; thus B increases with pres-sure, and the burning follows some logarithmic trend through ln (1þ B).

The form of Boq and Bfq presented in Eqn (6.120) is based on the assumption thatthe fuel droplet has infinite thermal conductivity, that is, the temperature of thedroplet is Ts throughout. But in an actual porous sphere experiment, the fuel entersthe center of the sphere at some temperature Tl and reaches Ts at the sphere surface.For a large sphere, the enthalpy required to raise the cool entering liquid to the sur-face temperature is cpl(Ts� Tl) where cpl is the specific heat of the liquid fuel. Toobtain an estimate of B that gives a conservative (lower) result of the burning ratefor this type of condition, one could replace Lv by:

L0v ¼ Lv þ cpl�Ts � Tl

�(6.122)

in the forms of B represented by Eqn (6.120).Table 6.1, extracted from Kanury [20], lists various values of B for many

condensed-phase combustible substances burning in air. Examination of this tablereveals that the variation of B for different combustible liquids is not great; rarely

Table 6.1 Transfer Numbers of Various Liquids in Aira

B B0 B00 B000

n-Pentane 8.94 8.15 8.19 9.00

n-Hexane 8.76 6.70 6.82 8.83

n-Heptane 8.56 5.82 6.00 8.84

n-Octane 8.59 5.24 5.46 8.97

i-Octane 9.43 5.56 5.82 9.84

n-Decane 8.40 4.34 4.62 8.95

n-Dodecane 8.40 4.00 4.30 9.05

Octene 9.33 5.64 5.86 9.72

Benzene 7.47 6.05 6.18 7.65

Methanol 2.95 2.70 2.74 3.00

Ethanol 3.79 3.25 3.34 3.88

Gasoline 9.03 4.98 5.25 9.55

Kerosene 9.78 3.86 4.26 10.80

Light diesel 10.39 3.96 4.40 11.50

Medium diesel 11.18 3.94 4.38 12.45

Heavy diesel 11.60 3.91 4.40 13.00

Acetone 6.70 5.10 5.19 6.16

Toluene 8.59 6.06 6.30 8.92

Xylene 9.05 5.76 6.04 9.48

Note: B ¼ ½imoNHþ cpðTN � TsÞ�=Lv; ¼ B0½imoNHþ cpðTN � TsÞ�=L0v;B00 ¼ ðimoNH=L0vÞ; B

000 ¼ ðimoNH=LvÞa TN¼ 20 �C.

340 CHAPTER 6 Diffusion flames

Page 41: Combustion || Diffusion flames

does one liquid fuel have a value of B a factor of 2 greater than another. Since thetransfer number always enters the burning rate expression as a ln(1þ B) term,one may conclude that, as long as the oxidizing atmosphere is kept the same, neitherthe burning rate nor the evaporation coefficient of liquid fuels will vary greatly. Thediffusivities and gas density dominate the burning rate. Whereas a tenfold variationin B results in an approximately twofold variation in burning rate, a tenfold variationof the diffusivity or gas density results in a tenfold variation in burning.

Furthermore, note that the burning rate has been determined without determiningthe flame temperature or the position of the flame. In the approach attributed to God-save [21] and extended by others [22,23], it was necessary to find the flame temper-ature, and the early burning rate developments largely followed this procedure. Theearly literature contains frequent comparisons not only of the calculated and exper-imental burning rates (or b) but also of the flame temperature and position. To theirsurprise, most experimenters found good agreement with respect to burning rate butpoorer agreement in flame temperature and position. What they failed to realize isthat, as shown by the discussion here, the burning rate is independent of the flametemperature. As long as an integrated approach is used and the gradients of temper-ature and product concentration are zero at the outer boundary, it does not matterwhat the reactions are or where they take place, provided they take place withinthe boundaries of the integration.

It is possible to determine the flame temperature Tf and position rf correspondingto the Godsave-type calculations simply by assuming that the flame exists at the po-sition where imo¼mf. Equation (6.85) is written in terms of bfo as:

r2s rsysrsDsr

¼ ln

�mfN � mfs � iðmoN � mosÞ þ ðmfs � 1Þ � imos

mf � mfs � iðmo � mosÞ þ ðmfs � 1Þ � imos

�(6.123)

At the flame surface, mf¼mo¼ 0 and mfN¼mos¼ 0; thus Eqn (6.123)becomes:

r2s rsysrsDsrf

¼ ln�1þ imoN

�(6.124)

Since _m ¼ 4pr2s rsys is known, rf can be estimated. Combining Eqns (6.118) and(6.124) results in the ratio of the radii:

rfrs

¼ lnð1þ BÞlnð1þ imoNÞ (6.125)

For the case of benzene (C6H6) burning in air, the mass stoichiometric index i isfound to be:

i ¼ 78

7:5� 32¼ 0:325

Since the mass fraction of oxygen in air is 0.23 and B0 from Table 6.1 is givenas 6, one has:

rfrs

¼ lnð1þ 6Þln½1þ ð0:325� 0:23Þ�y 27

6.3 Burning of Condensed Phases 341

Page 42: Combustion || Diffusion flames

This value is much larger than the value of about 2e4 observed experimen-tally. The large deviation between the estimated value and that observed ismost likely due to the assumptions made with respect to the thermophysical prop-erties and the Lewis number. This point is discussed in Section 6.3.2.3. Althoughthis estimate does not appear suitable, it is necessary to emphasize that the resultsobtained for the burning rate and the combustion evaporation coefficient derivedlater give good comparisons with experimental data. The reason for this supposedanomaly is that the burning rate is obtained by integration between the infinite at-mosphere and the droplet and is essentially independent of the thermophysicalparameters necessary to estimate internal properties. Such parameters includethe flame radius and the flame temperature, to be discussed in subsequentparagraphs.

For surface burning, as in an idealized case of char burning, it is appropriate toassume that the flame is at the particle surface and that rf¼ rs. Thus from Eqn(6.125), B must equal imoN. Actually, the same result would arise from the firsttransfer number in Eqn (6.120):

Bfo ¼ imoN þ mfs

1� mfs

Under the condition of surface burning, mfs¼ 0 and thus, again, Bfo¼ imoN.As in the preceding approach, an estimate of the flame temperature Tf at rf can be

obtained by writing Eqn (6.85) with b¼ boq. For the condition that Eqn (6.85) isdetermined at r¼ rf, one makes use of Eqn (6.124) to obtain the expression:

1þ imoN ¼�bN � bs þ 1

bf � bs þ 1

¼�HimoN þ cp

�TN � Ts

�þ Lv��

cp�Tf � Ts

�þ Lv� (6.126)

or:

cp�Tf � Ts

� ¼ ��HimoN þ cpðTN � TsÞ þ Lv�

1þ imoN

�� Lv

Again, comparisons of results obtained by Eqn (6.126) with experimental mea-surements show the same extent of deviation as that obtained for the determinationof rf/rs from Eqn (6.125). The reasons for this deviation are also the same.

Irrespective of these deviations, it is also possible from this transfer numberapproach to obtain some idea of the species profiles in the droplet burning case. Itis best to establish the conditions for the nitrogen profile for this quasi-steady fueldroplet burning case in air first. The conservation equation for the inert i (nitrogen)in droplet burning is:

4pr2rymi ¼ 4pr2rDdmi

dr(6.127)

342 CHAPTER 6 Diffusion flames

Page 43: Combustion || Diffusion flames

Integrating for the boundary condition that as r/N, mi¼miN, one obtains:

�4pr2s rsys4p

$1

r¼ rD ln

mi

miN

(6.128)

From either Eqn (6.86) or Eqn (6.118) it can be seen that:

rsrsys ¼ rsDs lnð1þ BÞThen Eqn (6.128) can be written in the form:

�rsrsysrsr¼ �½rsDs lnð1þ BÞ� rs

r¼ rD ln

mi

miN

(6.129)

For the condition at r¼ rs, this last expression reveals that:

�ln�1þ B

� ¼ ln

mis

miN

or:

1þ B ¼ miN

mis

Since the mass fraction of nitrogen in air is 0.77, for the condition By 6:

mis ¼ 0:77

7¼ 0:11

Thus the mass fraction of the inert nitrogen at the droplet surface is 0.11.Evaluating Eqn (6.129) at r¼ rf, one obtains:

�rsrflnð1þ BÞ ¼ ln

mif

miN

However, Eqn (6.125) gives:

rsrflnð1þ BÞ ¼ lnð1þ imoNÞ

Thus:

1þ imoN ¼ miN

mif

or for the case under consideration:

mif y0:77

1:075y 0:72

The same approach would hold for the surface burning of a carbon char exceptthat i would equal 0.75 (see Chapter 9) and:

mis ¼ 0:77

1:075� 0:22y 0:66

6.3 Burning of Condensed Phases 343

Page 44: Combustion || Diffusion flames

which indicates that the mass fraction of gaseous products mps for this carbon caseequals 0.34.

For the case of a benzene droplet in which B is taken as 6, the expression thatpermits the determination of mfs is:

B ¼ imoN þ mfs

1� mfs¼ 6

or:

mfs ¼ 0:85

This transfer number approach for a benzene droplet burning in air reveals spe-cies profiles characteristic of most hydrocarbons and gives the results summarizedbelow:

mfs ¼ 0:85 mff ¼ 0:00 mfN ¼ 0:00

mos ¼ 0:00 mof ¼ 0:00 mfN ¼ 0:23

mis ¼ 0:11 mif ¼ 0:72 miN ¼ 0:77

mps ¼ 0:04 mpf ¼ 0:28 mpN ¼ 0:00

These results are graphically represented in Figure 6.11. The product composi-tion in each case is determined for the condition that the mass fractions at each po-sition must total to 1.

It is now possible to calculate the burning rate of a droplet under the quasi-steadyconditions outlined and to estimate, as well, the flame temperature and position;however, the only means to estimate the burning time of an actual droplet is to calcu-late the evaporation coefficient for burning, b. From the mass burning resultsobtained, b may be readily determined. For a liquid droplet, the relation:

dm

dt¼ � _mf ¼ 4prlr

2s

drsdt

(6.130)

gives the mass rate in terms of the rate of change of radius with time. Here, rl is thedensity of the liquid fuel. It should be evaluated at Ts.

Many experimenters have obtained results similar to those given in Figure 6.13.These results confirm that the variation of droplet diameter during burning followsthe same “law” as that for evaporation:

d2 ¼ d20 � bt (6.131)

Since d¼ 2rs:

drsdt

¼ � b

8rs(6.132)

It is readily shown that Eqns (6.118) and (6.130) verify that a “d2” law shouldexist. Equation (6.130) is rewritten as:

_mf ¼ �2prlrs

dr2sdt

¼ �

prlrs2

��d�d2�dt

344 CHAPTER 6 Diffusion flames

Page 45: Combustion || Diffusion flames

Making use of Eqn (6.118):

_mf

4prs¼l

cp

lnð1þ BÞ ¼ �

rl8

��d�d2�dt

�¼ þ

rl8

�b

shows that:

d�d2�

dt¼ const; b ¼

8

rl

l

cp

lnð1þ BÞ (6.133)

which is a convenient form since l/cp is temperature-insensitive. Sometimes b iswritten as:

b ¼ 8

rs

rl

a ln

�1þ B

� ¼ 8

rD

rl

ln�1þ B

�to correspond more closely to expressions given by Eqns (6.96) and (6.118). Theproper response to Problem 9 of this chapter will reveal that the “d2” law is not appli-cable in modest Reynolds number or convective flow.

6.3.2.3 Refinements of the mass burning rate expressionSome major assumptions were made in the derivation of Eqns (6.118) and (6.133).First, it was assumed that the specific heat, thermal conductivity, and the product rDwere constants and that the Lewis number was equal to 1. Second, it was assumedthat there was no transient heating of the droplet. Furthermore, the role of finite re-action kinetics was not addressed adequately. These points will be given consider-ation in this section.

Variation of Transport Properties The transport properties used throughoutthe previous developments are normally strong functions of both temperature andspecies concentration. The temperatures in a droplet diffusion flame, as depictedin Figure 6.11, vary from a few hundred degrees Kelvin to a few thousand. In theregions of the droplet flame one essentially has fuel, nitrogen, and products. How-ever, under steady-state conditions as just shown, the major constituent in the sf

1.00.500.010

0.015

0.020

0.025

Benzene

t (sec)

d2 (cm2)

dropdiam

2

FIGURE 6.13

Diameteretime measurements of a benzene droplet burning in quiescent air showing

diameter-squared dependence.

After Godsave [21].

6.3 Burning of Condensed Phases 345

Page 46: Combustion || Diffusion flames

region is the fuel. It is most appropriate to use the specific heat of the fuel and itsbinary diffusion coefficient. In the region fN, the constituents are oxygen,nitrogen, and products. With similar reasoning, one essentially considers theproperties of nitrogen to be dominant; moreover, as the properties of fuel are usedin the sf region, the properties of nitrogen are appropriate to use in the fN region.To illustrate the importance of variable properties, Law [24] presented a simplifiedmodel in which the temperature dependence was suppressed, but the concentrationdependence was represented by using l, cp, and rD with constant, but different,values in the sf and fN regions. The burning rate result of this model was:

_m

4prs¼ ln

("1þ

�cp�sfðTf � TsÞL0v

#½l�ðcpÞsf�½1þ imoN�ðrDÞfN

)(6.134)

where Tf is obtained from expressions similar to Eqn (6.126) and found to be:

�cp�sf

�Tf � Ts

�þ �cp�fN

ðTf � TNÞhð1þ imoNÞ1=ðLeÞfN � 1

i ¼ H � L0v (6.135)

(rf/rs) was found to be:

rfrs

¼ 1þ"�

l�cp�sf

ðrDÞfNln�1þ �cp�sfðTf � TsÞ

�L0v�

lnð1þ imoNÞ

#

Law [24] points out that since imoN is generally much less than 1, the denomi-nator of the second term in Eqn (6.135) becomes [imoN/(Le)fN], which indicates thatthe effect of (Le)fN is to change the oxygen concentration by a factor ðLeÞ�1

fN asexperienced by the flame. Obviously, then, for (Le)fN> 1, the oxidizer concentra-tion is effectively reduced and the flame temperature is also reduced from the adia-batic value obtained for Le¼ 1, the value given by Eqn (6.126). The effective Lewisnumber for the mass burning rate (Eqn (6.134)) is:

Le ¼l

cp

sf

1

rD

fN

which reveals that the individual Lewis numbers in the two regions play no role indetermining _m. Law and Law [25] estimated that for heptane burning in air the effec-tive Lewis number was between 1/3 and 1/2. Such values in Eqn (6.136) predictvalues of rf/rs in the right range of experimental values [24].

The question remains as to the temperature at which to evaluate the physicalproperties discussed. One could use an arithmetic mean for the two regions [26]or Sparrow’s 1/3 rule [27], which gives:

Tsf ¼ 1

3Ts þ 2

3Tf ; TfN ¼ 1

3Tf þ 2

3TN

346 CHAPTER 6 Diffusion flames

Page 47: Combustion || Diffusion flames

Transient Heating of Droplets When a cold liquid fuel droplet is injected into ahot stream or ignited by some other source, it must be heated to its steady-state tem-perature Ts, derived in the last section. Since the heat-up time can influence the “d2”law, particularly for high-boiling-point fuels, it is of interest to examine the effectof the droplet heating mode on the main bulk combustion characteristicdtheburning time.

For this case, the boundary condition corresponding to Eqn (6.102) becomes:

_mLv þ4pr2ll

vT

vr

s�

¼4pr2lg

vT

vr

¼ _mL0v (6.136)

where the subscript s� designates the liquid side of the liquidesurfaceegas interfaceand sþ designates the gas side. The subscripts l and g refer to liquid and gas,respectively.

At the initiation of combustion, the heat-up (second) term of Eqn (6.136) can besubstantially larger than the vaporization (first) term. Throughout combustion thethird term is fixed. Thus, some [28,29] have postulated that droplet combustioncan be considered to consist of two periods: namely, an initial droplet heating periodof slow vaporization with:

_mL0vz4pr2ll

vT

vr

s�

followed by fast vaporization with almost no droplet heating so that:

_mL0vz _mLv

The extent of the droplet heating time depends on the mode of ignitionand the fuel volatility. If a spark is used, the droplet heating time can be areasonable percentage (10e20%) of the total burning time [27]. Indeed, theburning time would be 10e20% longer if the value of B used to calculate b iscalculated with L0v ¼ LV, that is, on the basis of the latent heat of vaporization.If the droplet is ignited by injection into a heated gas atmosphere, a long heatingtime will precede ignition; and after ignition, the droplet will be near its saturationtemperature so the heat-up time after ignition contributes little to the total burn-uptime.

To study the effects due to droplet heating, one must determine the temperaturedistribution T(r, t) within the droplet. In the absence of any internal motion, the un-steady heat transfer process within the droplet is simply described by the heat con-duction equation and its boundary conditions.

vT

vt¼ al

r2v

vr

r2vT

vr

; Tðr; t ¼ 0

� ¼ TlðrÞ;vT

vr

r¼0

¼ 0 (6.137)

The solution of Eqn (6.137) must be combined with the nonsteady equations forthe diffusion of heat and mass. This system can only be solved numerically and the

6.3 Burning of Condensed Phases 347

Page 48: Combustion || Diffusion flames

computing time is substantial. Therefore, a simpler alternative model of dropletheating is adopted [27,28]. In this model, the droplet temperature is assumed tobe spatially uniform at Ts and temporally varying. With this assumption, Eqn(6.136) becomes:

_mL0v ¼4pr2s lg

vT

vr

¼ _mLv þ4

3

pr3s rlcpl

vTsvt

(6.138)

But since:

_m ¼ �d

dt

�4

3

pr3s rl

�Equation (6.138) integrates to:

rsrso

3

¼ exp

8><>:� cpl

ZTs

Tso

dTs

L0v � Lv

9>=>; (6.139)

where L0v ¼ L0vðTsÞ is given by:

L0v ¼�1� mfs

��imoNH þ cp

�TN � Ts

��ðmfs þ imoNÞ (6.140)

and rso is the original droplet radius.Figure 6.14, taken from Law [29], is a plot of the nondimensional radius squared

as a function of a nondimensional time for an octane droplet initially at 300 Kburning under ambient conditions. Shown in this figure are the droplet historiescalculated using Eqns (6.137) and (6.138). Sirignano and Law [28] refer to the resultof Eqn (6.137) as the diffusion limit and that of Eqn (6.138) as the distillation limit,respectively. Equation (6.137) allows for diffusion of heat into the droplet, whereasEqn (6.138) essentially assumes infinite thermal conductivity of the liquid and hasvaporization at Ts as a function of time. Thus, one should expect Eqn (6.137) togive a slower burning time.

Also plotted in Figure 6.14 are the results from using Eqn (6.133) in whichthe transfer number was calculated with L0v ¼ Lv and L0v ¼ Lv þ cplðTs � TlÞ(Eqn (6.122)). As one would expect, L0v ¼ Lv gives the shortest burning time, andEqn (6.137) gives the longest burning time, since it does not allow storage of heatin the droplet as it burns. All four results are remarkably close for octane. The spreadcould be greater for a heavier fuel. For a conservative estimate of the burning time,use of B with L0v evaluated by Eqn (6.122) is recommended.

The Effect of Finite Reaction Rates When the fuel and oxidizer react at a finiterate, the flame front can no longer be considered infinitely thin. The reaction rateis then such that oxidizer and fuel can diffuse through each other and the reactionzone is spread over some distance. However, one must realize that although the

348 CHAPTER 6 Diffusion flames

Page 49: Combustion || Diffusion flames

reaction rates are considered finite, the characteristic time for the reaction is alsoconsidered to be much shorter than the characteristic time for the diffusional pro-cesses, particularly the diffusion of heat from the reaction zone to the droplet surface.

The development of the mass burning rate (Eqn (6.118)) in terms of the transfernumber B (Eqn (6.120)) was made with the assumption that no oxygen reaches thefuel surface and no fuel reaches N, the ambient atmosphere. In essence, the onlyassumption made was that the chemical reactions in the gas-phase flame zonewere fast enough so that the conditions mos¼ 0¼mfN could be met. The beautyof the transfer number approach, given that the kinetics are finite but faster thandiffusion and the Lewis number is equal to 1, is its great simplicity compared toother endeavors [21,22].

For infinitely fast kinetics, then, the temperature profiles form a discontinuity atthe infinitely thin reaction zone (see Figure 6.11). Realizing that the mass burningrate must remain the same for either infinite or finite reaction rates, one mustconsider three aspects dictated by physical insight when the kinetics are finite: first,the temperature gradient at r¼ rs must be the same in both cases; second, themaximum temperature reached when the kinetics are finite must be less than thatfor the infinite kinetics case; third, if the temperature is lower in the finite case,the maximum must be closer to the droplet in order to satisfy the first aspect. Lorellet al. [23] have shown analytically that these physical insights, as depicted inFigure 6.15, are correct.

00 0.1 0.2

0.2

0.4

0.6

0.8

1.0

(R/R

O)2

(α�/RO2) t

d2 —Law model for B

d2 —Law model for B1

Diffusion—limit modelRapid—mixing model

FIGURE 6.14

Variation of nondimensional droplet radius as a function of nondimensional time during

burning with droplet heating and steady-state models.

After Law [29].

6.3 Burning of Condensed Phases 349

Page 50: Combustion || Diffusion flames

6.4 BURNING OF DROPLET CLOUDSCurrent understanding of how particle clouds and sprays burn is still limited, despitenumerous studiesdboth analytical and experimentaldof burning droplet arrays.The main consideration in most studies has been the effect of droplet separationon the overall burning rate. It is questionable whether study of simple arrays willyield much insight into the burning of particle clouds or sprays.

An interesting approach to the spray problem has been suggested by Chiuand Liu [30], who consider a quasi-steady vaporization and diffusion processwith infinite reaction kinetics. They show the importance of a group combustionnumber (G), which is derived from extensive mathematical analyses and takesthe form:

G ¼ 3 1þ 0:276Re1=2Sc1=2Le N2=3

�RS

(6.141)

where Re, Sc, and Le are the Reynolds, Schmidt, and Lewis numbers, respectively; Nthe total number of droplets in the cloud; R the instantaneous average radius; and Sthe average spacing of the droplets.

The value of G was shown to have a profound effect upon the flame location anddistribution of temperature, fuel vapor, and oxygen. Four types of behaviors werefound for different ranges of G. External sheath combustion occurs for the largestvalue; and asG is decreased, there is external group combustion, internal group com-bustion, and isolated droplet combustion.

Isolated droplet combustion obviously is the condition for a separate flameenvelope for each droplet. Typically, a group number less that 10�2 is required.Internal group combustion involves a core with a cloud where vaporization existssuch that the core is totally surrounded by flame. This condition occurs for G

0.1440.1400.1360.1320.12800

0.2

0.4

0.6

1.2

0.8

1.0T3

T2T1 Z2

mO3

mO2

mF1mF2

mF3

mO1

Z1

0.32

0.31

0.30

Dim

ensi

onle

sste

mpe

ratu

re (

T°)

Position r (cm)

0.06

0.05

0.04

0.03

0.02

0.01

Wei

ght f

ract

ions

of

mf a

nd o

xidi

zer

mO

Fra

ctio

n of

rea

ctio

n co

mpl

eted

z

FIGURE 6.15

Effect of chemical rate processes on the structure of a diffusion-controlled droplet flame.

After Lorell et al. [23].

350 CHAPTER 6 Diffusion flames

Page 51: Combustion || Diffusion flames

values above 10�2 and somewhere below 1. As G increases, the size of the coreincreases. When a single flame envelops all droplets, external group combustionexists. This phenomenon begins with G values close to unity. (Note that many in-dustrial burners and most gas turbine combustors are in this range.) With externalgroup combustion, the vaporization of individual droplets increases with distancefrom the center of the core. At very high G values (above 102), only droplets in athin layer at the edge of the cloud vaporize. This regime is called the externalsheath condition.

6.5 BURNING IN CONVECTIVE ATMOSPHERES6.5.1 THE STAGNANT FILM CASEThe spherical-symmetric fuel droplet burning problem is the only quiescent case thatis mathematically tractable. However, the equations for mass burning may be readilysolved in one-dimensional form for what may be considered the stagnant film case.If the stagnant film is of thickness d, the free-stream conditions are thought to exist atsome distance d from the fuel surface (see Figure 6.16).

Within the stagnant film, the energy equation can be written as:

rycp

dT

dy

¼ l

d2T

dy2

þ _H (6.142)

With b defined as before, the solution of this equation and case proceeds as fol-lows. Analogous to Eqn (6.117), for the one-dimensional case:

ry

db

dy

¼ rD

d2b

dy2

(6.143)

Integrating Eqn (6.143), one has:

ryb ¼ rDdb

dyþ const (6.144)

mO,∞

y

Flame

Condensed phase fuel

FIGURE 6.16

Stagnant film height representation for condensed-phase burning.

6.5 Burning in Convective Atmospheres 351

Page 52: Combustion || Diffusion flames

The boundary condition is:

rD

db

dy

o

¼ rsys ¼ ry (6.145)

Substituting this boundary condition into Eqn (6.144), one obtains:

rybo ¼ ryþ const; ryðbo � 1Þ ¼ const

The integrated equation now becomes:

ry�b� bo þ 1

� ¼ rDdb

dy(6.146)

which upon second integration becomes:

ryy ¼ rD lnðb� bo þ 1Þ þ const (6.147)

At y¼ 0, b¼ bo; therefore, the constant equals 0, so that:

ryy ¼ rD lnðb� bo þ 1Þ (6.148)

Since d is the stagnant film thickness:

ryd ¼ rD lnðbd � bo þ 1Þ (6.149)

Thus:

Gf ¼rD

d

ln�1þ B

�(6.150)

where, as before:

B ¼ b d � bo or b N � bo (6.151)

Since the Prandtl number cpm/l is equal to 1, Eqn (6.150) may be written as:

Gf ¼rD

d

lnð1þ BÞ ¼

l

cpd

lnð1þ BÞ ¼

md

�lnð1þ BÞ (6.152)

A burning pool of liquid or a volatile solid fuel will establish a stagnant filmheight due to the natural convection that ensues. From analogies to heat transferwithout mass transfer, a first approximation to the liquid pool burning rate may bewritten as:

Gfd

m lnð1þ BÞwGra (6.153)

where a equals 1/4 for laminar conditions and 1/3 for turbulent conditions, d is thecritical dimension of the pool, and Gr is the Grashof number:

Gr ¼gd3b1a2

DT (6.154)

where g is the gravitational constant and b1 is the coefficient of expansion.

352 CHAPTER 6 Diffusion flames

Page 53: Combustion || Diffusion flames

When the free streamdbe it forced or due to buoyancy effectsdis transverse tothe mass evolution from the regressing fuel surface, no stagnant film forms, in whichcase the correlation given by Eqn (6.153) is not explicitly correct.

6.5.2 THE LONGITUDINALLY BURNING SURFACEMany practical cases of burning in convective atmospheres may be approximated byconsidering a burning longitudinal surface. This problem is similar to what could becalled the burning flat plate case and has application to the problems that arise in thehybrid rocket. It differs from the stagnant film case in that it involves gradients in thex direction as well as the y direction. This configuration is depicted in Figure 6.17.

For the case in which the Schmidt number is equal to 1, it can be shown [7] thatthe conservation equations (in terms of U; see Eqn (6.17)) can be transposed into theform used for the momentum equation for the boundary layer. Indeed, the transfor-mations are of the same form as the incompressible boundary layer equations devel-oped and solved by Blasius [31]. The important difference between this problem andthe Blasius [31] problem is the boundary condition at the surface. The Blasius equa-tion takes the form:

f000 þ ff 00 ¼ 0 (6.155)

where f is a modified stream function and the primes designate differentiation withrespect to a transformed coordinate. The boundary conditions at the surface for theBlasius problem are:

h ¼ 0; f ð0Þ ¼ 0; and f 0ð0Þ ¼ 0

h ¼ N; f 0ðNÞ ¼ 1(6.156)

For the mass burning problem, the boundary conditions at the surface are:

h ¼ 0; f 0ð0Þ ¼ 0; and Bf 00ð0Þ ¼ 0

h ¼ N; f 0ðNÞ ¼ 1(6.157)

where h is the transformed distance normal to the plate. The second of these condi-tions contains the transfer number B and is of the same form as the boundary con-dition in the stagnant film case (Eqn (6.145)).

Condensed phase fuel

Oxidizer

y

Flame

x

FIGURE 6.17

Burning of a flat fuel surface in a one-dimensional flow field.

6.5 Burning in Convective Atmospheres 353

Page 54: Combustion || Diffusion flames

Emmons [32] solved this burning problem, and his results can be shown to be ofthe form:

Gf ¼l

cp

Re

1=2x

xffiffiffi2

p!½ � f ð0Þ� (6.158)

where Rex is the Reynolds number based on the distance x from the leading edge ofthe flat plate. For Prandtl number equal to 1, Eqn (6.158) can be written in the form:

Gfx

m¼ Re

1=2x ½ � f ð0Þ�ffiffiffi

2p (6.159)

It is particularly important to note that [�f(0)] is a function of the transfernumber B. This dependence as determined by Emmons is shown in Figure 6.18.

An interesting approximation to this result (Eqn (6.159)) can be made from thestagnant film result of the last section, that is:

Gf ¼rD

d

lnð1þ BÞ ¼

l

cpd

lnð1þ BÞ ¼

md

�lnð1þ BÞ (6.152)

If d is assumed to be the Blasius boundary layer thickness dx, then:

dx ¼ 5:2xRe�1=2x (6.160)

Substituting Eqn (6.160) into Eqn (6.152) gives:

Gfx

m¼ Re

1=2x

5:2

!lnð1þ BÞ (6.161)

The values predicted by Eqn (6.161) are somewhat high compared to those pre-dicted by Eqn (6.159). If Eqn (6.161) is divided by B0.15/2 to give:

Gfx=m ¼ Re

1=2x

2:6

!�lnð1þ BÞB0:15

�(6.162)

1.0

B

00

8

1 10 100

4

2

6

In(1�B )B0.152.6

x �

[�f(

0)]

FIGURE 6.18

[f(0)] as a function of the transfer number B.

354 CHAPTER 6 Diffusion flames

Page 55: Combustion || Diffusion flames

the agreement is very good over a wide range of B values. To show the extent of theagreement, the function:

lnð1þ BÞB0:152:6

(6.163)

is plotted on Figure 6.18 as well.Obviously, these results do not hold at very low Reynolds numbers. As Re

approaches zero, the boundary layer thickness approaches infinity. However, theburning rate is bounded by the quiescent results.

6.5.3 THE FLOWING DROPLET CASEWhen droplets are not at rest relative to the oxidizing atmosphere, the quiescentresults no longer hold so forced convection must again be considered. No onehas analytically solved this complex case. As discussed in Chapter 4, Section4.8, flow around a sphere (or cylinder as previously illustrated) can be complex,and at relatively high Re(>20), there is a boundary layer flow around the frontof the sphere and a wake or eddy region behind it.

For this burning droplet case, an overall heat transfer relationship could be writ-ten to define the boundary condition given by Eqn (6.90):

hðDTÞ ¼ GfLv (6.164)

The thermal driving force is represented by a temperature gradient DT, which isthe ambient temperature TN plus the rise in this temperature due to the energyrelease minus the temperature at the surface Ts, or:

DT ¼ TN þimoNH

cp

� Ts ¼

�imoNH þ cpðTN � TsÞ

cp

�(6.165)

Substituting Eqns (6.165) and (6.118) for Gf into Eqn (6.164), one obtains:

h�imoNH þ cpðTN � TsÞ

�cp

¼�

rD

r

lnð1þ BÞ

�Lv

¼�

l

cpr

lnð1þ BÞ

�Lv

(6.166)

where r is now the radius of the droplet. Upon cross-multiplication, Eqn (6.166)becomes:

hr

l¼ lnð1þ BÞ

B¼ Nu (6.167)

since:

B ¼�imoNH þ cpðTN � TsÞ

Lv

6.5 Burning in Convective Atmospheres 355

Page 56: Combustion || Diffusion flames

Since Eqn (6.118) was used for Gf, this Nusselt number (Eqn (6.167)) is forthe quiescent case (Re/ 0). For small B, ln(1þ B)z B and the Nu¼ 1, the clas-sical result for heat transfer without mass addition.

The term (ln(1þ B))/B has been used as an empirical correction for higher Rey-nolds number problems, and a classical expression for Nu with mass transfer is:

Nur ¼�lnð1þ BÞ

B

� 1þ 0:39Pr1=3Re1=2r

�(6.168)

As Re/ 0, Eqn (6.168) approaches Eqn (6.167). For the case Pr¼ 1 andRe> 1, Eqn (6.168) becomes:

Nur ¼ ð0:39Þ�lnð1þ BÞ

B

�Re1=2r (6.169)

The flat plate result of the preceding section could have been written in terms of aNusselt number as well. In that case:

Nux ¼h� f ð0ÞRe1=2x

iffiffiffi2

pB

(6.170)

Thus, the burning rate expressions related to Eqns (6.169) and (6.170) are,respectively:

Gfr=m ¼ 0:39Re1=2r lnð1þ BÞ (6.171)

¼ Re1=2r ½�f ð0Þ�ffiffiffi

2p (6.172)

In convective flow a wake exists behind the droplet and droplet heat transfer inthe wake may be minimal, so these equations are not likely to predict quantitativeresults. Note, however, that if the right-hand side of Eqn (6.171) is divided byB0.15, the expressions given by Eqns (6.171) and (6.172) follow identical trends;thus data can be correlated as:

Gfr

m

�B0:15

lnð1þ BÞ�versus Re1=2r (6.173)

If turbulent boundary layer conditions are achieved under certain conditions, thesame type of expression should hold, and Re should be raised to the 0.8 power.

If, indeed, Eqns (6.171) and (6.172) adequately predict the burning rate ofa droplet in laminar convective flow, the droplet will follow a “d3/2” burning ratelaw for a given relative velocity between the gas and the droplet. In this caseb will be a function of the relative velocity as well as B and other physical pa-rameters of the system. This result should be compared to the “d2” law(Eqn (6.172)) for droplet burning in quiescent atmospheres. In turbulent flow,droplets will appear to follow a burning rate law in which the power of thediameter is close to 1.

356 CHAPTER 6 Diffusion flames

Page 57: Combustion || Diffusion flames

6.5.4 BURNING RATES OF PLASTICS: THE SMALL B ASSUMPTIONAND RADIATION EFFECTS

Current concern with the fire safety of polymeric (plastic) materials has promptedgreat interest in determining the mass burning rate of plastics. For plastics whoseburning rates are measured so that no melting occurs, or for nonmelting plastics,the developments just obtained should hold. For the burning of some plastics inair or at low oxygen concentrations, the transfer number may be considered smallcompared to 1. For this condition, which of course would hold for any system inwhich B<< 1:

lnð1þ BÞyB (6.174)

and the mass burning rate expression for the case of nontransverse air movementmay be written as:

Gf y

l

cpd

B (6.175)

Recall that for these considerations the most convenient expression for B is:

B ¼�imoNH þ cpðTN � TsÞ

�Lv

(6.176)

In most cases:

imoNH > cpðTN � TsÞ (6.177)

so:

ByimoNH

Lv(6.178)

and:

Gf y

l

cpd

imoNH

Lv

(6.179)

Equation (6.179) shows why good straighteline correlations are obtained whenGf is plotted as a function of moN for burning rate experiments in which the dy-namics of the air are constant or well controlled (i.e., d is known or constant).One should realize that:

GfwmoN (6.180)

holds only for small B.The consequence of this small B assumption may not be immediately apparent.

One may obtain a physical interpretation by again writing the mass burning rateexpression for the two assumptions made (i.e., B<< 1 and B¼ [imoN H]/Lv):

Gf y

l

cpd

imoNH

Lv

(6.181)

6.5 Burning in Convective Atmospheres 357

Page 58: Combustion || Diffusion flames

and realizing that as an approximation:

ðimoNHÞy cpðTf � TsÞ (6.182)

where Tf is the diffusion flame temperature. Thus, the burning rate expressionbecomes:

Gf y

l

cpd

�cpðTf � TsÞ

Lv

�(6.183)

Cross-multiplying, one obtains:

GfLvylðTf � TsÞ

d(6.184)

which says that the energy required to gasify the surface at a given rate perunit area is supplied by the heat flux established by the flame. Equation(6.184) is simply another form of Eqn (6.164). Thus, the significance of the smallB assumption is that the gasification from the surface is so small that it does notalter the gaseous temperature gradient determining the heat flux to the surface.This result is different from the result obtained earlier, which stated that the stag-nant film thickness is not affected by the surface gasification rate. The small Bassumption goes one step further, revealing that under this condition the tempera-ture profile in the boundary layer is not affected by the small amount ofgasification.

If flame radiation occurs in the mass burning processdor any other radi-ation is imposed, as is frequently the case in plastic flammability testsdonecan obtain a convenient expression for the mass burning rate provided oneassumes that only the gasifying surface, and none of the gases between theradiation source and the surface, absorbs radiation. In this case Fineman[33] showed that the stagnant film expression for the burning rate can be approx-imated by:

Gf ¼

l

cpd

ln

�1þ

B

1� E

�where Eh

QR

GfLv

and QR is the radiative heat transfer flux.This simple form for the burning rate expression is possible because the equa-

tions are developed for the conditions in the gas phase and the mass burning ratearises explicitly in the boundary condition to the problem. Since the assumptionis made that no radiation is absorbed by the gases, the radiation term appears onlyin the boundary condition to the problem.

Notice that as the radiant flux QR increases, E and the term B/(1� E) increase.When E¼ 1, the problem disintegrates because the equation was developed in theframework of a diffusion analysis. E¼ 1 means that the solid is gasified by theradiant flux alone.

358 CHAPTER 6 Diffusion flames

Page 59: Combustion || Diffusion flames

PROBLEMS1. A laminar fuel jet issues from a tube into air and is ignited to form a flame.

The height of the flame is 8 cm. With the same fuel the diameter of the jet isincreased by 50% and the laminar velocity leaving the jet is decreased by 50%.What is the height of the flame after the changes are made? Suppose theexperiments are repeated but that grids are inserted in the fuel tube so that allflows are turbulent. Again for the initial turbulent condition it is assumed theflame height is 8 cm.

2. An ethylene oxide monopropellant rocket motor is considered part of aram rocket power plant in which the turbulent exhaust of the rocket reactswith induced air in an afterburner. The exit area of the rocket motor is 8 cm2.After testing it is found that the afterburner length must be reduced by 42.3%.What size must the exit port of the new rocket be to accomplish this reductionwith the same afterburner combustion efficiency? The new rocket wouldoperate at the same chamber pressure and area ratio. How many of the newrockets would be required to maintain the same level of thrust as the originalpower plant?

3. A spray of benzene fuel is burned in quiescent air at 1 atm and 298 K.The burning of the spray can be assumed to be characterized by single dropletburning. The (Sauter) mean diameter of the spray is 100 mm, that is, the burningtime of the spray is the same as that of a single droplet of 100 mm. Calculate amean burning time for the spray. For calculation purposes, assume whatevermean properties of the fuel, air, and product mixture are required. For someproperties those of nitrogen would generally suffice. Also assume that the dropletis essentially of infinite conductivity. Report, as well, the steady-state tempera-ture of the fuel droplet as it is being consumed.

4. Repeat the calculation of the previous problem for the initial condition that theair is at an initial temperature of 1000 K. Further calculate the burning time forthe benzene in pure oxygen at 298 K. Repeat all calculations with ethanol as thefuel. Then discuss the dependence of the results obtained on ambient conditionsand fuel type.

5. Two liquid sprays are evaluated in a single-cylinder test engine. The first is liquidmethanol, which has a transfer number B¼ 2.9. The second is a regular dieselfuel, which has a transfer number B¼ 3.9. The two fuels have approximately thesame liquid density; however, the other physical characteristics of the dieselspray are such that its Sauter mean diameter is 1.5 times that of the methanol.Both are burning in air. Which spray requires the longer burning time, and howmuch longer is it than the other?

6. Consider each of the condensed-phase fuels listed to be a spherical particleburning with a perfect spherical flame front in air. From the properties of thefuels given, estimate the order of the fuels with respect to mass burning rate perunit area. List the fastest burning fuel first, etc.

Problems 359

Page 60: Combustion || Diffusion flames

7. Suppose fuel droplets of various sizes are formed at one end of a combustor andmove with the gas through the combustor at a velocity of 30 m/s. It is known thatthe 50-mm droplets completely vaporize in 5 ms. It is desired to vaporizecompletely each droplet of 100 mm and less before they exit the combustionchamber. What is the minimum length of the combustion chamber allowable indesign to achieve this goal?

8. A radiative flux QR is imposed on a solid fuel burning in air in a stagnation filmmode. The expression for the burning rate is:

Gf ¼rD

d

ln

�1þ

b

1� E

�where E¼QR/GfLs. Develop this expression. It is a one-dimensional problem asdescribed.

9. Experimental evidence from a porous sphere burning rate measurement in alow Reynolds number laminar flow condition confirms that the mass burning rateper unit area can be represented by:

Gfr

m

¼�f 0ð0; BÞffiffiffi

2p

�Re1=2r

Would a real droplet of the same fuel follow a “d2” law under the same condi-tions? If not, what type of power law should it follow?

10. Write what would appear to be the most important physical result in each of thefollowing areas:a. Laminar flame propagationb. Laminar diffusion flamesc. Turbulent diffusion flamesd. Detonatione. Droplet diffusion flames

Explain the physical significance of the answers. Do not develop equations.

Latent

heat of

vaporization

(cal/gm)

Density

(gm/cm3)

Thermal

conductivity

(cal/s-cm-K)

Stoichiometric

heat evolution

in air per unit

weight of fuel

(cal/gm)

Heat

capacity

(cal/gm/K)

Aluminum 2570 2.7 0.48 1392 0.28

Methanol 263 0.8 0.51� 10�3 792 0.53

Octane 87 0.7 0.33� 10�3 775 0.51

Sulfur 420 2.1 0.60� 10�3 515 0.25

360 CHAPTER 6 Diffusion flames

Page 61: Combustion || Diffusion flames

11. Composition measurements in a propane air diffusion flame indicate the majorspecies as C3H8, O2, N2, CO, H2, CO2 and H2O. Define the mixture fraction interms of the various species mole fractions. Determine the numerical value forthe stoichiometric mixture fraction.

REFERENCES[1] Gaydon AG, Wolfhard H. Flames. New York: Macmillan; 1960 [chapter 6].[2] Smyth KC, Miller JH, Dorfman RC, Mallard WG, Santoro RJ. Combust Flame 1985;

62:157.[3] Kang KT, Hwang JY, Chung SH, Lee W. Combust Flame 1997;109:266.[4] Du J, Axelbaum RL. Combust Flame 1995;100:367.[5] Hottel HC, Hawthorne WR. Proc Combust Inst 1949;3:255.[6] Boedecker LR, Dobbs GM. Combust Sci Technol 1986;46:301.[7] Williams FA. Combustion theory. 2nd ed. Menlo Park (California): Benjamin-

Cummins; 1985 [chapter 3].[8] Lewis B, von Elbe G. Combustion, flames and explosion of gases. 2nd ed. New York:

Academic Press; 1961.[9] Burke SP, Schumann TEW. Ind Eng Chem 1928;20:998.[10] Roper FG. Combust Flame 1977;29:219.[11] Roper FG, Smith C, Cunningham AC. Combust Flame 1977;29:227.[12] Flowers WL, Bowman CT. Proc Combust Inst 1986;21:1115.[13] Bilger RW. Combust Sci Technol 1976;13:155.[14] Linan A, Williams FA. Fundamental aspects of combustion. Oxford (England): Oxford

University Press; 1995.[15] Hawthorne WR, Weddel DB, Hottel HC. Proc Combust Inst 1949;3:266.[16] Khudyakov L. Chem Abstr 1955;46:10844e.[17] Blackshear Jr PL. An introduction to combustion. Minneapolis: Department of

Mechanical Engineering, University of Minnesota; 1960 [chapter 4].[18] Spalding DB. Proc Combust Inst 1953;4:846.[19] Spalding DB. Some fundamentals of combustion. London: Butterworth; 1955 [chapter 4].[20] Kanury AM. Introduction to combustion phenomena. New York: Gordon & Breach;

1975 [chapter 5].[21] Godsave GAE. Proc Combust Inst 1953;4:818.[22] Goldsmith M, Penner SS. Jet Propul 1954;24:245.[23] Lorell J, Wise H, Carr RE. J Chem Phys 1956;25:325.[24] Law CK. Prog Energy Combust Sci 1982;8:171.[25] Law CK, Law AV. Combust Sci Technol 1976;12:207.[26] Law CK, Williams FA. Combust Flame 1972;19:393.[27] Hubbard GL, Denny VE, Mills AF. Int J Heat Mass Transf 1975;18:1003.[28] Sirignano WA, Law CK. Adv Chem Ser No 1978;166:1.[29] Law CK. Combust Flame 1976;26:17.[30] Chiu HH, Liu TM. Combust Sci Technol 1977;17:127.[31] Blasius H. Z Math Phys 1956;56:1.[32] Emmons HW. Z Angew Math Mech 1956;36:60.[33] Fineman S. Some analytical considerations of the hybrid rocket combustion problem

[M.S.E. thesis]. Princeton (NJ): Department of Aeronautical Engineering; 1962.

References 361